Loading [MathJax]/jax/element/mml/optable/SpacingModLetters.js
Research article

Decay properties for evolution-parabolic coupled systems related to thermoelastic plate equations

  • In this paper, we consider the Cauchy problem for a family of evolution-parabolic coupled systems, which are related to the classical thermoelastic plate equations containing non-local operators. By using diagonalization procedure and WKB analysis, we derive representation of solutions in the phase space. Then, sharp decay properties in a framework of LpLq are investigated via these representations. Particularly, some thresholds for the regularity-loss type decay properties are found.

    Citation: Zihan Cai, Yan Liu, Baiping Ouyang. Decay properties for evolution-parabolic coupled systems related to thermoelastic plate equations[J]. AIMS Mathematics, 2022, 7(1): 260-275. doi: 10.3934/math.2022017

    Related Papers:

    [1] Jianxia He, Ming Li . Existence of global solution to 3D density-dependent incompressible Navier-Stokes equations. AIMS Mathematics, 2024, 9(3): 7728-7750. doi: 10.3934/math.2024375
    [2] Kaile Chen, Yunyun Liang, Nengqiu Zhang . Global existence of strong solutions to compressible Navier-Stokes-Korteweg equations with external potential force. AIMS Mathematics, 2023, 8(11): 27712-27724. doi: 10.3934/math.20231418
    [3] Yasir Nadeem Anjam . The qualitative analysis of solution of the Stokes and Navier-Stokes system in non-smooth domains with weighted Sobolev spaces. AIMS Mathematics, 2021, 6(6): 5647-5674. doi: 10.3934/math.2021334
    [4] Xiaoxia Wang, Jinping Jiang . The pullback attractor for the 2D g-Navier-Stokes equation with nonlinear damping and time delay. AIMS Mathematics, 2023, 8(11): 26650-26664. doi: 10.3934/math.20231363
    [5] Yasir Nadeem Anjam . Singularities and regularity of stationary Stokes and Navier-Stokes equations on polygonal domains and their treatments. AIMS Mathematics, 2020, 5(1): 440-466. doi: 10.3934/math.2020030
    [6] José Luis Díaz Palencia, Saeed ur Rahman, Antonio Naranjo Redondo . Analysis of travelling wave solutions for Eyring-Powell fluid formulated with a degenerate diffusivity and a Darcy-Forchheimer law. AIMS Mathematics, 2022, 7(8): 15212-15233. doi: 10.3934/math.2022834
    [7] Qingkun Xiao, Jianzhu Sun, Tong Tang . Uniform regularity of the isentropic Navier-Stokes-Maxwell system. AIMS Mathematics, 2022, 7(4): 6694-6701. doi: 10.3934/math.2022373
    [8] Hui Fang, Yihan Fan, Yanping Zhou . Energy equality for the compressible Navier-Stokes-Korteweg equations. AIMS Mathematics, 2022, 7(4): 5808-5820. doi: 10.3934/math.2022321
    [9] Qasim Khan, Anthony Suen, Hassan Khan, Poom Kumam . Comparative analysis of fractional dynamical systems with various operators. AIMS Mathematics, 2023, 8(6): 13943-13983. doi: 10.3934/math.2023714
    [10] Krunal B. Kachhia, Jyotindra C. Prajapati . Generalized iterative method for the solution of linear and nonlinear fractional differential equations with composite fractional derivative operator. AIMS Mathematics, 2020, 5(4): 2888-2898. doi: 10.3934/math.2020186
  • In this paper, we consider the Cauchy problem for a family of evolution-parabolic coupled systems, which are related to the classical thermoelastic plate equations containing non-local operators. By using diagonalization procedure and WKB analysis, we derive representation of solutions in the phase space. Then, sharp decay properties in a framework of LpLq are investigated via these representations. Particularly, some thresholds for the regularity-loss type decay properties are found.



    Since Engle's groundbreaking work on the univariate ARCH model [12], and the subsequent generalization to the GARCH model by Bollerslev [9], as well as the introduction of the threshold GARCH by Glosten et al. [20], these models have become fundamental tools for capturing the key stylized facts of financial time series. The capability to model the co-movements of multiple series is of paramount practical importance. Analyzing them as components of a multivariate process is of particular benefit, especially in the presence of temporal or contemporaneous dependencies. The spectrum of multivariate (A)GARCH models is expanding, with a substantial body of literature already established [3,15,26,28]. This includes significant probabilistic, structural, and asymptotic findings that support various estimation methods for a wide range of multivariate GARCH models and their extensions.

    Engle and Kroner [13] set the groundwork for covariance stationarity in multivariate generalized ARCH models. The existence of a 4th-moment structure in the CCC-GARCH model was validated by He and Terasvirta [22] and Aue et al. [1]. Jeantheau [23] demonstrated the robust consistency of the Quasi-Maximum Likelihood Estimator (QMLE) in multivariate ARCH models. Ling and McAleer [25] furthered this by establishing the asymptotic properties of the QMLE for vector ARMA-GARCH models, contingent on the finiteness of the 4th-moment of the observed process. Notably, while asymptotic results for some asymmetric models were known, the works of [2] extended these to a broader class of multidimensional causal processes, including asymmetries. McAleer et al. [27] contributed to the structure and asymptotic theory of the VARMA-AGARCH model, a direction further expanded by Francq and Zakoïan [14,15], who provided results for a variety of multivariate (A)GARCH models without relying on moment assumptions for the observed process. Extensions of the EGARCH model to the multivariate context were also proposed by Koutmos and Booth [24]. In terms of more recent developments, stationary multivariate GARCH models have evolved from time-invariant to time-varying coefficients. These models allow coefficients to depend on an unobservable, time-homogeneous Markov chain, as demonstrated in the multivariate MS-GARCH model by Haas and Liu [21].

    Francq and Zakoïan (2010, [14]) investigated the quasi-maximum likelihood estimation of a class of symmetric multivariate GARCH models without imposing moment conditions for the observed process. They extended this study to multivariate asymmetric GARCH models in their subsequent paper (Francq and Zakoïan, 2012, [15]). In [5], Bibi (2018) extended the symmetric multivariate GARCH models studied by Francq and Zakoïan (2010, [14]) by investigating the asymptotic properties of QML estimation for symmetric multivariate periodic CCC-GARCH models. Maïnassara et al. (2022, [26]) made additions that can be considered an extension of Francq and Zakoïan's models by studying the estimation of multivariate asymmetric power GARCH models.

    We extend the work undertaken in these studies. On the one hand, we extend the standard models by introducing coefficients related to time-varying and periodicity (for more details, see Ghezal et al. [16,17]); on the other, we extend Bibi's paper by incorporating the effect of asymmetry. More accurately, in our paper, we expanded the standard models to include the effects of temporal and periodic factors on the model coefficients. Additionally, we incorporate the effects of asymmetry in multivariate GARCH models, allowing for cross-leverage effects, which was not considered in Bibi's paper, which focused solely on symmetric models. Moreover, our paper illustrates our major findings by application to a bivariate exchange rates series, while Bibi's paper focused solely on the theoretical aspects of periodic multivariate CCC-GARCH models. Furthermore, our primary contribution lies in providing asymptotic results for this model category without imposing moment assumptions for the observed process, a topic further explored by Ghezal et al. [7,16]. These differences highlight the unique contribution of our research to the development of multivariate GARCH models.

    In this context, the following symbols are utilized:

    I(n) represents the n×n identity matrix and O(k,l) represents a matrix of size k×l with all entries being zeros. For clarity, let us consider: O_(k):=O(k,1) and O(k):=O(k,k).

    As signifies AA...A repeated s times, where represents the usual Kronecker product of matrices. Vec(A) denotes the vector stacking operator. If (A(k),kK) is a sequence of n×n matrices, it is important to note that for any integers m and t, ti=mA(i)=A(m)A(m+1)A(t) if mt, and I(n) otherwise.

    ρ(A) represents the spectral radius of the squared matrix A, det(A) (resp. Tr(A)) denotes the determinant (resp. trace) of A. Moreover, for any matrices A, B, and C with suitable dimensions, Tr(BA)=Vec(B)Vec(A) and Vec(BAC)=(CB)Vec(A). Let . represent any induced matrix norm on the set of r×p and r×1 matrices. This norm verifies, for any p×r matrix A and r×p matrix B, that |Tr(AB)|nm AB, A2Tr(AA)mA2, and |AA|A2m.

    θ_ represents the vector of first-order partial derivatives with respect to θ_, 2θ_ represents the matrix of second-order partial derivatives with respect to θ_, and represents convergence in the distribution.

    Recall that if g(B) is a real-valued function of a matrix B depending on some parameter θ_, then θ_g(B)=Tr(g(B)Bθ_B). Furthermore, assuming the invertibility of matrix A, the following holds true for any vector x_ and matrices B and C with suitable dimensions:

    θ_A1=A1(θ_A)A1

    and

    x_Ax_A=x_x_,log|det(A)|A=A1,Tr(BAC)A=CB,Tr(BA1C)A=A1CBA1,Tr(BACA)A=BAC+CAB.

    We begin with the CCC-PAGARCH models defined for a rdimensional {X_t=(X1t,...,Xrt)} as follows

    {X_t=V1/2tη_t,Var{X_t|Ft1}=Vt=HtMHt, Ht=diag(σ_1/2t),σ_t=a_0(t)+qk=1(Ak(t)X_+tk+Bk(t)X_tk)+pj=1Cj(t)σ_tj., (2.1)

    where σ_τt:=(στ1t,...,στrt), τR+, X_+t:=(X+21t,...,X+2rt), X_t:=(X21t,...,X2rt), z=(z)+=(z)0, z±2=(z±)2, Ft:=σ{X_ti,i0}, M represents a constant conditional correlation matrix, a_0(.) represents a r×1 vector whose entries are strictly positive, the Ai(.), Bi(.) and Cj(.) for all i,j are r×r matrices with positive entries, the functions a_0(t),Ai(t), Bi(t) and Cj(t) exhibit periodicity in t with a period of s, and (η_t) is an i.i.d. random vector with a mean and covariance matrix (O_(r),I(r)). Now, it is important to emphasize that this model is limited to the periodic case, i.e., by setting t=ns+v, where v=1,...,s and nZ. Consequently, the model (2.1) can be expressed in the following form

    {X_ns+v=V1/2ns+vη_ns+v,Vns+v=Hns+vMHns+v, Hns+v=diag(σ_1/2ns+v),σ_ns+v=a_0(v)+qk=1(Ak(v)X_+ns+vk+Bk(v)X_ns+vk)+pj=1Cj(v)σ_ns+vj.. (2.2)

    CCC-PAGARCH is a comprehensive framework that encompasses several models previously explored in the literature. This includes standard CCC-AGARCH models when s=1, as discussed in [15], conventional symmetric CCC-PeriodicGARCH models when A.(.)=B.(.), as explored in [5], and periodic asymmetric GARCH models when r=1, as studied by Ghezal et al. [6]. Now, such as seen in several time series models, it is usually beneficial to express (2.1) as an analogous stochastic recurrence equation to facilitate a more streamlined analysis (see [17,18,19]). For this reason, we write X_t=Ht~η_t, X_+t=G+tσ_t and X_t=Gtσ_t, where ~η_t=M1/2η_t and G±t= diag(~η_t±2), and we consider the r(2q+p)×r(2q+p) matrix

    Δt:=(¯(G+Aq)t(G+qBq)t(G+qCp)t(GqCq)t¯(GAq)t(GqCp)t(IpAq)t(IpBq)t¯(IAp)t),

    where

    (Φ+nΨm)t=(Φ+tΨ1(t)Φ+tΨm(t)O(r(n1),rm)),¯(Φ+Ψm)t=(Φ+tΨ1(t)Φ+tΨm(t)I(r(m1))O(r(m1),r)).

    Consequently, we have X_t=Gtσ_t and σ_t=FΛ_t, such that

    Λ_t=ΔtΛ_t1+Π_t, (2.3)

    where

    Λ_t:=((X_+t),(X_+tq+1),(X_t),,(X_tq+1),σ_t,,σ_tp+1),
    F:=(O(r,2rq)I(r)O(r,r(p1))),

    and

    Π_t:=(a_0(t)G+t,O_(r(q1)),a_0(t)Gt,O_(r(q1)),a_0(t),O_(r(p1))).

    Now, iterating (2.3) stimes, we get

    Γ_t:=Λ_st=ΩtΓ_t1+Ξ_t, (2.4)

    in which Ωt=s1j=0Δstj and Ξ_t=s1i=0 {i1l=0Δstl}Π_sti where, as is customary, emptied products are defined to be equal to I(r(2q+p)). Note here that (Ωt) (resp. (Ξ_t)) represents a sequence of i.i.d. positive random matrices (resp. vectors), independent of Γ_n, n<t. Therefore, a causal, periodically ergodic, and strictly periodical solution (short form: PE and SPS) for Eq (2.3), with an equivalent representation in (2.2), holds true if Eq (2.4) has a causal, ergodic, and strictly stationary (short form: SS) solution. Observe that according to (2.3) and (2.4) are correct for all integer t, through recursive substitution, we derive formal series solutions denoted, respectively, by

    Γ_t(1)=i0{i1l=0Δtl}Π_ti

    and

    Γ_t(2)=i0{i1l=0Ωtj}Ξ_ti.

    The uniqueness of the non-anticipative SS solution (Γ_t) to the model (2.4) appears in a form comparable to that of Bougerol and Picard [10] through the use of the topLyapunov exponent γΩ linked to the sequence of stochastic matrices Ω=(Ωt) defined by

    γΩ:=infm1{1mE{logΩmΩm1...Ω1}}=limm1mlogΩmΩm1...Ω1 a.s.

    Furthermore, since E{log+Ω1} E{Ω1}<+ where log+z represents logz0, z>0, this ensures the existence of γΩ. Building upon the preceding discussion, we are ready to unveil the following outcomes.

    Theorem 1. Consider the process (X_t)tZ defined by the model (2.1) with a periodic vector representation (2.3), and suppose that E{η_t}<+. The following statements are thus equivalent:

    The Eq (2.4) having a unique, strictly stationary, causal and ergodic solution provided by Γ_t(2).

    The Eq (2.3) having a unique, SPS, causal and PE solution provided by Γ_t(1).

    γΩ<0.

    Furthermore, for all tZ, the series Γ_t(1) and Γ_t(2) converge absolutely almost surely, Γ_t(1) a.s=Γ_t(2).

    We also need the subsequent simple outcome which characterizes a necessary condition for SS.

    Corollary 1. If the CCCPAGARCH(p,q) model described in (2.1) satisfies γΩ<0, thus ρ(s1v=0Csv)<1, where Ct=(C1(t)Cp(t)I(r(p1))O(r(p1),r)).

    Now, the subsequent proposition introduces conditions that are slightly more robust, guaranteeing the negativity of γΩ.

    Proposition 1. Consider the CCCPAGARCH(p,q) model, then

    a. ρ(E{Ω1})<1 implies that γΩ<0.

    b. E{logt1i=0Ωti}<0 or E{t1i=0Ωti}<1 for some t1 γΩ<0.

    Moreover, if γΩ<0, then there is ς>0 such that t, E{σ_tς}<+, E{X_tς}<+.

    The following results establish the geometric ergodicity and βmixing of Markov chain (Γ_t) with state space Rr(2q+p) whose nstep transition probability is given by Pnx_0(D)=P(Γ_tD|Γ_0=x_0) for any Borel set DBRr(2q+p).

    Theorem 2. If (X_t)tZ is the process governed by (2.1), characterized by a state-space representation as described in (2.4), then

    i. ρ(E{Ω1})<1 entails the existence of only one invariant distribution of Px_(.), x_, a non-negative Borel measurable function h1 defined on Rr(2q+p), a compact set W in BRr(2q+p), a constant κ and δ]0,1[ such that for any x_Wc, E{h(Γ_n+1)|Γ_n=x}δh(x_)κ and supx_WE{h(Γ_n+1)|Γ_n=x}<+.

    ii. The chain (Γ_n) is geometrically ergodic and βmixing property with exponential decay rate with E{Γ_n}<+ if, and only if, ρ(E{Ω1})<1.

    iii. sup1vsρ(qi=1(Ai(v)+Bi(v))+pj=1Cj(v))<1 implies that the chain (Γ_n) is geometrically ergodic.

    In this section, we consider the QMLE for estimating the parameters of the CCCPAGARCH(p,q) model assembled in

    θ_=(θ1,...,θl):=(θ_1,...,θ_s,m_)Θ(]0;+[r×[0;+[r2(p+2q))s×]1;1[r(r1)/2

    where

    l=s˜r+r(r1)/2,θ_v:=(θ1(v),...,θ˜r(v),1vs):=(a_0(v),α_(v),β_(v),γ_(v))

    and

    ˜r=r+r2(2q+p) 

    with

    α_(v):=(Vec(A1(v)),...,Vec(Aq(v))),β_(v):=(Vec(B1(v)),...,Vec(Bq(v))),γ_(v):=(Vec(C1(v)),...,Vec(Cp(v))),

    for all 1vs, θi(v)=a0,i(v),i=1,...,r, θr+ijk(v)=αijk(v),i,j=1,...,r,k=1,...,q, θr+r2q+ijk(v)=βijk(v),i,j=1,...,r,k=1,...,q, θr+2r2q+ijk(v)=γijk(v),i,j=1,...,r,k=1,...p and M=(mij) is the symmetric matrix with mii=1 (i.e., m_:=(m21,...,mr1,m32,...,mr2,...,mrr1)). The true parameter value, symbolized by θ_0 and belonging to the parameter space Θ, is unknown and needs to be estimated. In pursuit of this, consider that the observations X_1,...,X_n=sN constitute a time series from the unique, SPS and causal solution of (2.1). A QMLE of θ_ is defined as any measurable solution ˆθ_n of

    ˆθ_n=Argmaxθ_Θ˜LNs(θ_)=Argminθ_Θ(˜LNs(θ_)).

    For initial values X_0,...,X_1q,˜σ_0(θ_),...,˜σ_1p(θ_), the Gaussian loglikelihood function for θ_Θ is given up to an additive constant by

    ˜LNs(θ_)=(Ns)1Nt=1s1v=0˜lst+v(θ_),

    with ˜lt(θ_)=logdet(˜Vt)+X_t˜V1tX_t, where ˜Vt is recursively defined for t1 by

    {˜Vns+v=˜Hns+vM˜Hns+v, ˜Hns+v=diag(˜σ_1/2ns+v),˜σ_ns+v=a_0(v)+qi=1(Ai(v)X_+ns+vi+Bi(v)X_ns+vi)+pj=1Cj(v)˜σ_ns+vj,.

    Due to the strong dependency of ˜Vt(θ_) on initial values, the process (˜lt(θ_)) is neither an S PS nor PE process. Hence, it would be appropriate to replace the process (˜lt(θ_)) with its SP S and PE version. Therefore, we introduce an approximate version LNs(θ_)=(Ns)1Nt=1s1v=0lst+v(θ_) of the likelihood (4.2) with lt(θ_)=logdet(Vt)+X_tV1tX_t. The upcoming findings in this paper confirm the robust consistency of ˆθ_Ns and its asymptotic normality.

    Take into account the regularity assumptions listed below:

    A0 θ_0Θ and Θ is compact.

    A1 Consider the polynomials A0,v(z)=qi=1A0,i(v)zi, B0,v(z)=qi=1B0,i(v)zi and C0,v(z)=I(r)pj=1C0,j(v)zj adheres to the convention A0,v(z)=B0,v(z)=O(r) if q=0 and C0,v(z)=I(r) if p=0, for all v{1,...,s}. If p>0, A0,v(1)+B0,v(1)O, A0,v(z), B0,v(z) and C0,v(z) are left-coprime and [A0,q(v) B0,q(v)  C0,p(v)] has full rank r for all v{1,...,s}.

    A2 γΩ0<0 with Ω0 in place of Ω to confirm that the unknown parameter is θ_0 and ρ(s1v=0C0,sv)<1.

    A3 The distribution of the components of ˜η_t isn't focused on two points and P(˜ηit(0,+))(0,1) for i=1,,r. A4 M represents a correlation matrix that is positive-definite.

    Currently, we have the capability to articulate our initial finding.

    Theorem 3. Assuming the specified Assumptions, A0–A4, ˆθ_Ns exhibits robust consistency, meaning that

    P(ˆθ_NsNθ_0)=1.

    To establish the proof of Theorem 3, we articulate the subsequent technical statements encapsulated in the following lemma.

    Lemma 1. Assuming the specified Assumptions, A0–A4, we obtain

    Ⅰ. supθ_Θ|(˜LNsLNs)(θ_)|a.s.N0.

    Ⅱ. There exists an integer tZ such that σ_t(θ_)a.s.=σ_t(θ_0) and M=M0 θ_=θ_0.

    Ⅲ. sv=1Eθ_0{lst+v(θ_0)}< and sv=1Eθ_0{lst+v(θ_)} is minimized at θ_=θ_0.

    Ⅳ. θ_θ_0, there exists a neighborhood V(θ_) such that sv=1Eθ_0{lst+v(θ_0)}<liminfNinf˜θ_Θ(˜LNs(˜θ_)) a.s.

    To demonstrate the AN of ˆθ_Ns, the analysis relies on additional assumptions.

    A5 θ_0Θ˚ where \mathring{\Theta} represents the interior of the parameter space \Theta.

    \textbf{A6} E\left\{ \left\Vert \underline{\eta}_{t}\underline{\eta }_{t}^{\prime}\right\Vert ^{2}\right\} < \infty.

    These supplementary assumptions are essential to establishing the required conditions under which the estimator \widehat{\underline{\theta} }_{Ns} exhibits asymptotic normality. The second principal outcome is encapsulated in the following theorem

    Theorem 4. Assuming that \left(\underline{X}_{t}, t\in\mathbb{Z}\right) is formed by the model referenced in (2.1), under the conditions specified in A0–A6 we obtain

    \left( Ns\right) ^{\frac{1}{2}}\left( \widehat{\underline{\theta}} _{Ns}-\underline{\theta}_{0}\right) \leadsto\mathcal{N}\left( \underline {O},K^{-1}UK^{-1}\right) \text{ as }N\rightarrow \infty,

    where the matrix K = \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \nabla_{\underline{\theta}}^{2}l_{st+v}\left(\underline{\theta}_{0}\right) \right\} is positive-definite and the matrix

    U = \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \nabla_{\underline{\theta}}l_{st+v}\left( \underline{\theta}_{0}\right) \nabla_{\underline{\theta}}^{\prime}l_{st+v}\left( \underline{\theta} _{0}\right) \right\} .

    The demonstration of Theorem 4 relies on the classical technique of a Taylor series expansion of \nabla_{\underline{\theta}} L_{Ns}\left(\underline{\theta}\right) around \underline{\theta}_{0} , and this expansion is expressed as

    0 = \left( Ns\right) ^{-\frac{1}{2}}\sum\limits_{t = 1}^{Ns}\nabla _{\underline{\theta}}\widetilde{l}_{t}\left( \widehat{\underline{\theta} }_{Ns}\right) = \left( Ns\right) ^{-\frac{1}{2}}\sum\limits_{t = 1}^{Ns} \nabla_{\underline{\theta}}\widetilde{l}_{t}\left( \underline{\theta} _{0}\right) +\left( \left( Ns\right) ^{-1}\sum\limits_{t = 1}^{Ns} \nabla_{\underline{\theta}}^{2}\widetilde{l}_{t}\left( \widetilde {\underline{\theta}}\right) \right) \left( Ns\right) ^{\frac{1}{2}}\left( \widehat{\underline{\theta}}_{Ns}-\underline{\theta}_{0}\right) .

    Here, the coordinates of \widetilde{\underline{\theta}} lie between \widehat{\underline{\theta}}_{Ns} and \underline{\theta}_{0}. The theorem will consequently follow straightforwardly. To achieve this, we will establish the subsequent intermediate outcomes, consolidated in the following lemma

    Lemma 2. Assuming A0–A6, we obtain

    \boldsymbol{a.} \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \underset{\underline{\theta}\in\mathcal{\Theta}}{\sup}\left\Vert \nabla_{\underline{\theta}}l_{st+v}\left(\underline{\theta}_{0}\right) \nabla_{\underline{\theta}}^{\prime}l_{st+v}\left(\underline{\theta} _{0}\right) \right\Vert \right\} < \infty and \sum\limits_{v = 1} ^{s}E\left\{ \underset{\underline{\theta}\in\mathcal{\Theta}}{\sup}\left\Vert \nabla_{\underline{\theta}}^{2}l_{st+v}\left(\underline{\theta}_{0}\right) \right\Vert \right\} < \infty .

    \boldsymbol{b.} K is invertible and the existence of the matrix U.

    \boldsymbol{c.} \exists\mathcal{V}\left(\underline{\theta}_{0}\right) of \underline{\theta}_{0} such that \sum\nolimits_{v = 1}^{s}E_{\underline {\theta}_{0}}\left\{ \underset{\underline{\theta}\in\mathcal{V}\left(\underline{\theta}_{0}\right) }{\sup}\left\vert \dfrac{\partial^{3} l_{st+v}\left(\underline{\theta}\right) }{\partial\theta_{i}\left(v\right) \partial\theta_{j}\left(v\right) \partial\theta_{k}\left(v\right) }\right\vert \right\} < \infty for all 1\leq i, j, k\leq l.

    \textbf{d.} \left(Ns\right) ^{-\frac{1}{2}}\sum\limits_{t = 1} ^{Ns}\nabla_{\underline{\theta}}l_{t}\left(\underline{\theta}_{0}\right) \ \leadsto\mathcal{N}\left(\underline{O}, U\right) \ as N\rightarrow \infty.

    Remark 4.1. The extension of the existing models to accommodate asymmetries, as highlighted in the symmetric CCC-PGARCH paper by Bibi [5], addresses a crucial aspect of volatility modeling in financial time series data. Asymmetries arise when positive and negative past values of the process have different impacts on the current volatility, thereby affecting the current volatility differently. This phenomenon introduces complexities in characterizing the usual leverage effect, particularly when considering the possibility of positive or negative volatility values. To address these asymmetries, researchers have proposed extensions to existing models, such as asymmetric specifications of GARCH models, and incorporating correlations between volatility process noise and observation series noise in the stochastic volatility framework. The aim of these efforts is to enhance the models' ability to capture the nuanced dynamics of financial markets, including asymmetric responses to market shocks. Therefore, the proposed extension to the class of CCC-PGARCH models fills an important gap in the literature and contributes to advancing our understanding of volatility dynamics in financial time series data.

    Remark 4.2. In our study, we introduce the concept of an expanded model derived from our primary model, presenting an open problem that invites scientific inquiry and exploration. This extended model, termed the double-switching CCC- \left(A\right) GARCH model, combines the principles of a CCC- \left(A\right) GARCH model with Markov switching and a periodic sequence (see., Zerari et al. [30] for further ideas). The aim of integrating these elements is to establish a robust analytical framework to comprehend volatility within financial markets. By enhancing the predictive capabilities of volatility models and refining risk management strategies, this approach not only contributes to more effective financial analysis but also lays the groundwork for future research endeavors in the field of financial data analysis.

    Remark 4.3. The extension of Bibi's theories in our paper offers substantial value-added contributions to this field. While Bibi's work primarily focused on extending symmetric multivariate GARCH models to include periodic features and investigate the asymptotic properties of QML estimation, our paper goes further by incorporating the effects of asymmetry. This extension is crucial as it allows for a more comprehensive understanding of the dynamics of financial time series data, particularly in capturing cross-leverage effects and asymmetric behavior. Additionally, our paper provides asymptotic results for a class of generally nonstationary multivariate GARCH models, a topic that has not been extensively explored in the existing literature. By doing so without imposing moment assumptions on the observed process, we contribute to a deeper understanding of the underlying processes governing financial time series data, thereby enhancing the applicability and robustness of multivariate GARCH models in real-world scenarios.

    Herein, we present a simulation study to assess the effectiveness of the QMLE, where we simulated 500 independent trajectories, each with lengths of n = 1000 and n = 3000 , from a bivariate periodic stationary CCC- AGARCH(1.1) model, with \mathcal{N}(\underline{O}_{\left(2\right) }, I_{\left(2\right) }) and t_{15} as innovations distributions with a \underline{O}_{\left(2\right) } mean and a I_{\left(2\right) } covariance matrix. The parameter \underline{\theta}_{0} is selected to meet the SPS condition \gamma_{\Omega} < 0 . The Root Mean Square Errors (RMSE) of \widehat{{ \underline{\theta}}} \left(v\right) , where v = 1, \dots, s , are computed to assess the performance of the estimators.

    Our initial Monte Carlo study involves the estimation of a bivariate standard CCC-AGARCH(1.1) model, with the parameters associated with this model gathered in vector \underline{\theta} = \left(\underline {a}_{0}^{\prime}, Vec^{\prime}\left(A_{1}\right), Vec^{\prime}\left(B_{1}\right), Vec^{\prime}\left(C_{1}\right), m_{21}\right) ^{\prime}, where C_{1} is a diagonal matrix, implying that the volatility of each component is associated with its own lagged value, as well as the lagged values of the squared observations of both components. The results of the simulation, reported in Table 1, are in accordance with the consistency.

    Table 1.  Results of a simulation study for a bivariate standard CCC-AGARCH with a variety of sample sizes n and a number of replications equal to 500.
    \mathcal{N}(\underline{O}_{\left(2\right) }, I_{\left(2\right) }) t_{15}
    n 1000 3000 1000 3000
    True value Mean RMSE Mean RMSE Mean RMSE Mean RMSE
    \underline{a}_{0} 1.50 1.4998 0.0002 1.4999 0.0001 1.4997 0.0003 1.4999 0.0001
    1.00 0.9910 0.0090 0.9924 0.0076 0.9901 0.0099 0.9918 0.0082
    Vec\left(A_{1}\right) 0.15 0.1445 0.0055 0.1449 0.0051 0.1443 0.0057 0.1446 0.0054
    0.10 0.1014 0.0014 0.1010 0.0010 0.1020 0.0020 0.1012 0.0012
    0.05 0.0533 0.0033 0.0524 0.0024 0.0538 0.0038 0.0531 0.0031
    0.24 0.2434 0.0066 0.2432 0.0032 0.2437 0.0037 0.2433 0.0033
    Vec\left(B_{1}\right) 0.50 0.4978 0.0022 0.4983 0.0017 0.4976 0.0024 0.4979 0.0021
    0.07 0.0710 0.0010 0.0707 0.0007 0.0721 0.0021 0.0709 0.0009
    0.06 0.0618 0.0018 0.0610 0.0010 0.0671 0.0071 0.0625 0.0025
    0.20 0.2031 0.0031 0.2019 0.0019 0.2034 0.0034 0.2028 0.0028
    Vec\left(C_{1}\right) 0.10 0.0982 0.0018 0.0984 0.0016 0.0981 0.0019 0.0983 0.0017
    0.76 0.7663 0.0063 0.7616 0.0016 0.7686 0.0086 0.7648 0.0048
    m_{21} 0.80 0.8057 0.0057 0.8052 0.0052 0.8061 0.0061 0.8059 0.0059

     | Show Table
    DownLoad: CSV

    Our second Monte Carlo study involved the estimation of a bivariate periodic CCC-AGARCH(1.1) model, with the parameters associated with this model gathered in vector \underline{\theta} = \left(\underline {\theta}^{\prime}\left(1\right), \underline{\theta}^{\prime}\left(2\right), m_{21}\right) ^{\prime} with \underline{\theta}\left(v\right) = \left(\underline{a}_{0}^{\prime}\left(v\right), Vec^{\prime}\left(A_{1}\left(v\right) \right), Vec^{\prime}\left(B_{1}\left(v\right) \right), Vec^{\prime }\left(C_{1}\left(v\right) \right) \right) ^{\prime} , is chosen to ensure the SPS condition, where the matrix C_{1} (1) (resp. C_{1} (2)) is diagonal, implying that the volatility of each component is associated with its own lagged value, as well as the lagged values of the squared observations of both components. The first results of the simulation presented in Table 2 align with the consistency.

    Table 2.  Results of a simulation study for a bivariate periodic CCC-AGARCH (1.1) model, with a variety of sample sizes n and a number of replications equal to 500.
    \mathcal{N}(\underline{O}_{\left(2\right) }, I_{\left(2\right) }) t_{15}
    n 1000 3000 1000 3000
    \underline{\theta} True value Mean RMSE Mean RMSE Mean RMSE Mean RMSE
    \underline{a}_{0} (1) 1.50 1.4958 0.0042 1.4966 0.0034 1.4950 0.0050 1.4961 0.0039
    1.50 1.5016 0.0016 1.5012 0.0012 1.5018 0.0018 1.5014 0.0014
    Vec\left(A_{1}\left(1\right) \right) 0.50 0.4930 0.0070 0.4939 0.0061 0.4926 0.0074 0.4938 0.0062
    0.15 0.1407 0.0093 0.1415 0.0085 0.1402 0.0098 0.1414 0.0086
    0.15 0.1523 0.0023 0.1518 0.0018 0.1524 0.0024 0.1517 0.0017
    \underline{\theta} (1) 0.25 0.2535 0.0035 0.2528 0.0028 0.2537 0.0037 0.2528 0.0028
    Vec\left(B_{1}\left(1\right) \right) 0.20 0.1916 0.0084 0.1976 0.0024 0.1919 0.0081 0.1949 0.0051
    0.05 0.0030 0.0030 0.0013 0.0013 0.0063 0.0063 0.0018 0.0018
    0.05 0.0573 0.0073 0.0545 0.0045 0.0579 0.0079 0.0549 0.0049
    0.35 0.3557 0.0057 0.3523 0.0023 0.3563 0.0063 0.3528 0.0028
    Vec\left(C_{1}\left(1\right) \right) 0.10 0.1019 0.0019 0.1010 0.0010 0.1034 0.0034 0.1017 0.0017
    0.60 0.6440 0.0440 0.6378 0.0378 0.6485 0.0485 0.6391 0.0391
    \underline{a}_{0} (2) 1.50 1.4980 0.0020 1.4985 0.0015 1.4978 0.0022 1.4983 0.0017
    1.50 1.5013 0.0013 1.5011 0.0011 1.5016 0.0016 1.5013 0.0013
    Vec\left(A_{1}\left(2\right) \right) 0.45 0.4393 0.0107 0.4408 0.0092 0.4390 0.0110 0.4407 0.0093
    0.19 0.1833 0.0067 0.1848 0.0052 0.1820 0.0080 0.1847 0.0053
    0.20 0.2085 0.0085 0.2075 0.0075 0.2096 0.0096 0.2078 0.0078
    \underline{\theta} (2) 0.45 0.4534 0.0034 0.4517 0.0017 0.4548 0.0048 0.4521 0.0021
    Vec\left(B_{1}\left(2\right) \right) 0.25 0.2397 0.0103 0.2416 0.0016 0.2384 0.0116 0.2414 0.0014
    0.10 0.0968 0.0032 0.0996 0.0004 0.0951 0.0049 0.0996 0.0004
    0.10 0.1141 0.0141 0.1126 0.0126 0.1158 0.0158 0.1127 0.0127
    0.35 0.3564 0.0064 0.3542 0.0042 0.3588 0.0088 0.3545 0.0045
    Vec\left(C_{1}\left(2\right) \right) 0.40 0.4062 0.0062 0.4009 0.0009 0.4068 0.0068 0.4015 0.0015
    0.34 0.3478 0.0078 0.3460 0.0060 0.3483 0.0083 0.3473 0.0073
    m_{21} 0.80 0.8168 0.0168 0.8106 0.0106 0.8259 0.0259 0.8154 0.0154

     | Show Table
    DownLoad: CSV

    From Tables 1 and 2, and as expected, the QMLE estimates for the coefficients of the periodic CCC-AGARCH(1.1) model exhibit a decrease in RMSE as the sample size increases. Additionally, it is observed that the estimates based on the normal distribution \mathcal{N}\left(0, 1\right) generally demonstrate greater efficiency compared to those corresponding to the t- distribution with 15 degrees of freedom t_{15} .

    This section is devoted to modeling, with periodic CCC-AGARCH(1, 1) model, two datasets, namely the daily time series (X_{1t})_{t\geq1} for the exchange rate Euro/Algerian dinar (EUR/DZD) , and the daily time series (X_{2t})_{t\geq1} for the exchange rate U.S. dollar/ Algerian dinar (USD/DZD) .

    We first removed all the days when the market was closed, including holidays and weekends. The observations extended over the period from 2000-01-03 to 2011-09-29. In order to do so, we collected the two exchange rates at time t in the vector \underline{X}_{t}^{\prime} = \left(X_{1t}, X_{2t}\right) = \left(EUR/DZD_{t}, USD/DZD_{t}\right), and their corresponding \log- return series in the vector \underline{r}_{t}^{\prime} = \left(r_{1t}, r_{2t}\right), where r_{kt} = \log\left(X_{kt}/X_{kt-1}\right), for k = 1, 2. The plots of prices \left(X_{1t}, X_{2t}\right) and the daily returns series of prices \left(r_{1t}, r_{2t}\right), squad and absolute returns are plotted in Figures 1 and 2.

    Figure 1.  The plots of the price series \left(X_{1t}\right), \left(r_{1t}\right), \left(r_{1t}^{2}\right) and \left(\left\vert r_{1t}\right\vert \right) .
    Figure 2.  The plots of the price series \left(X_{2t}\right), \left(r_{2t}\right), \left(r_{2t}^{2}\right) and \left(\left\vert r_{2t}\right\vert \right) .

    The findings indicate that for the EUR/DZD (resp. USD/DZD ), the lowest returns ( -0.0233 ) (resp. -0.0191 ) and the highest returns ( 0.0497 ) (resp. 0.0335 ). The skewness for two \log- return series is positive. Moreover, one of the features which prominently stands out from Tables 3 and 4 is that the kurtosis for two \log- returns is much larger than 3, suggesting that models dependent on the Gaussian assumption may not adequately describe the data. Now, we suggest the 5-periodic CCC- AGARCH(1.1) model that permits for the description of the intraweek effect in the daily exchange rate, where the parameters are allowed to vary with the day of the week, v = 1 corresponds to Monday, v = 2 to Tuesday, and so forth. The estimated parameters of the 5 -periodic CCC- (A)GARCH(1.1) model and their RMSE are reported in Tables 5 and 6.

    Table 3.  Summary statistics for daily exchange rate series (X_{1t})_{t\geq1}, their returns, \left(r_{1t}\right)_{t\geq 1}, \left(\left\vert r_{1t}\right\vert \right)_{t\geq1} and \left(r_{1t}^{2}\right)_{t\geq1}.
    Series mean Std. Dev Median Skewness Kurtosis Min Max Arch(300) J. Bera LBtest
    \left(X_{1t}\right) 88.612 11.575 91.099 -0.5181 2.1329 67.204 109.07 100\% 2.324\times10^{2} 100\%
    \left(r_{1t}\right) 11.816\times10^{-5} 50.427\times10^{-4} 12.346\times10^{-5} 0.3536 8.9678 -0.0233 0.0497 100\% 4.597\times10^{3} 24.20\%
    \left(\left\vert r_{1t}\right\vert \right) 35.753\times10^{-4} 35.575\times10^{-4} 0.2554\times10^{-2} 2.6956 18.431 0.0000 0.0497 100\% 3.401\times10^{4} 100\%
    \left(r_{1t}^{2}\right) 2.5434\times10^{-5} 7.1925\times10^{-5} 6.523\times10^{-6} 16.103 464.37 0.0000 0.0025 30, 67\% 2.723\times10^{7} 100\%

     | Show Table
    DownLoad: CSV
    Table 4.  Summary statistics for daily exchange rate series \left(X_{2t}\right) _{t\geq1}, their returns, \left(r_{2t}\right) _{t\geq1}, \left(\left\vert r_{2t}\right\vert \right)_{t\geq1} and \left(r_{2t}^{2}\right) _{t\geq1}.
    Series mean Std. Dev Median Skewness Kurtosis Min Max Arch(300) J. Bera LBtest
    \left(X_{2t}\right) 73.451 4.2424 73.126 -0.6005 3.7642 60.345 81.282 100\% 2.580\times10^{2} 100\%
    \left(r_{2t}\right) 2.199\times10^{-5} 30.251\times10^{-4} -6.431\times10^{-6} 0.7162 13.007 -0.0191 0.0335 100\% 1.301\times10^{4} 43.29\%
    \left(\left\vert r_{2t}\right\vert \right) 18.78\times10^{-4} 23.713\times10^{-4} 10.606\times10^{-4} 3.0114 20.804 0.0000 0.0335 100\% 4.497\times10^{4} 100\%
    \left(r_{2t}^{2}\right) 9.149\times10^{-6} 3.173\times10^{-5} 1.1249\times10^{-6} 17.678 531.45 0.0000 0.0011 00\% 3.571\times10^{7} 100\%

     | Show Table
    DownLoad: CSV
    Table 5.  QMLE estimate and their RMSE for 5-periodic CCC-AGARCH(1, 1) model.
    Days \widehat{\underline{a}}_{0} \widehat{A}_{1} \widehat{B}_{1} \widehat{C}_{1} \widehat{m}_{21}
    Monday \left(\begin{array}{c} 0.6021 \\ (0.0001) \\ 1.0649 \\ (0.0012) \end{array}\right) \left(\begin{array}{ll} 0.0532 & 0.0071 \\ (0.0156) & (0.0092) \\ 0.0081 & 0.5563 \\ (0.0752) & (0.0154) \end{array}\right) \left(\begin{array}{ll} 0.0838 & 0.0001 \\ (0.0157) & (0.0000) \\ 0.5436 & 1.0154 \\ (0.0047) & (0.0026) \end{array}\right) {diag}\left(\left(\begin{array}{c} 0.8036 \\ (0.0000) \\ 0.0879 \\ (0.0583) \end{array}\right)^{\prime}\right) 0.6438
    (0.0325)
    Tuesday \left(\begin{array}{c} 2.8023 \\ (0.1725) \\ 0.5018 \\ (0.0034) \end{array}\right) \left(\begin{array}{ll} 0.0878 & 0.5136 \\ (0.0135) & 0.0468) \\ 0.0632 & 0.5051 \\ (0.0076) & (0.0085) \end{array}\right) \left(\begin{array}{ll} 0.1266 & 0.2072 \\ (0.0149) & 0.0185) \\ 0.7687 & 0.0327 \\ (0.0624) & (0.0619) \end{array}\right) { diag }\left(\left(\begin{array}{c} 0.4485 \\ (0.0285) \\ 0.1596 \\ (0.0094) \end{array}\right)^{\prime}\right)
    Wednesday \left(\begin{array}{c} 1.5257 \\ (0.2452) \\ 0.3766 \\ (0.0651) \end{array}\right) \left(\begin{array}{ll} 0.0713 & 0.0711 \\ (0.0138) & (0.0652) \\ 0.1133 & 0.5386 \\ (0.0215) & (0.1123) \end{array}\right) \left(\begin{array}{ll} 0.1264 & 0.1864 \\ (0.0096) & (0.0207) \\ 0.3381 & 0.6061 \\ (0.0234) & (0.0175) \end{array}\right) { diag }\left(\left(\begin{array}{c} 0.6931 \\ (0.0355) \\ 0.2743 \\ (0.0739) \end{array}\right)^{\prime}\right)
    Thursday \left(\begin{array}{c} 1.3665 \\ (0.5006) \\ 0.0000 \\ (0.0014) \end{array}\right) \left(\begin{array}{ll} 0.0506 & 0.2765 \\ (0.0153) & (0.0319) \\ 0.0597 & 0.4648 \\ (0.0079) & (0.1140) \end{array}\right) \left(\begin{array}{ll} 0.0672 & 0.1151 \\ (0.0114) & (0.0145) \\ 0.3216 & 0.9808 \\ (0.0233) & (0.0526) \end{array}\right) { diag }\left(\left(\begin{array}{c} 0.7134 \\ (0.0814) \\ 0.0391 \\ (0.0899) \end{array}\right)^{\prime}\right)
    Friday \left(\begin{array}{c} 0.4513 \\ (0.0000) \\ 0.2405 \\ (0.0084) \end{array}\right) \left(\begin{array}{ll} 0.0549 & 0.0611 \\ (0.0169) & (0.0528) \\ 0.0739 & 0.2680 \\ (0.0055) & (0.0438) \end{array}\right) \left(\begin{array}{ll} 0.1169 & 0.4327 \\ (0.0143) & 0.0352) \\ 0.5524 & 0.4693 \\ (0.0647) & (0.0462) \end{array}\right) { diag }\left(\left(\begin{array}{c} 0.9261 \\ (0.0000) \\ 0.2717 \\ (0.0452) \end{array}\right)^{\prime}\right)

     | Show Table
    DownLoad: CSV
    Table 6.  QMLE estimate and their RMSE for 5- periodic CCC- GARCH(1, 1) model.
    Days \widehat{\underline{a}}_{0} \widehat{A}_{1} \widehat{B}_{1} \widehat{C}_{1} \widehat{m}_{21}
    Monday \left(\begin{array}{c} 0.8139 \\ (0.0065) \\ 1.0927 \\ (0.0096) \end{array}\right) \left(\begin{array}{ll} 0.0758 & 0.0360 \\ (0.0173) & (0.0128) \\ 0.0213 & 0.5801 \\ (0.0778) & (0.0179) \end{array}\right) { diag }\left(\left(\begin{array}{c} 0.8158 \\ (0.0070) \\ 0.1005 \\ (0.0637) \end{array}\right)^{\prime}\right) 0.8162
    (0.0578)
    Tuesday \left(\begin{array}{c} 3.1603 \\ (0.2136) \\ 0.5601 \\ (0.0111) \end{array}\right) \left(\begin{array}{ll} 0.0905 & 0.5419 \\ (0.0153) & (0.0497) \\ 0.0684 & 0.5148 \\ (0.0148) & (0.0099) \end{array}\right) {diag}\left(\left(\begin{array}{c} 0.4042 \\ (0.0323) \\ 0.1388 \\ (0.0139) \end{array}\right)^{\prime}\right)
    Wednesday \left(\begin{array}{c} 1.7182 \\ (0.2907) \\ 0.4154 \\ (0.0749) \end{array}\right) \left(\begin{array}{ll} 0.0720 & 0.0843 \\ (0.0142) & (0.0810) \\ 0.1352 & 0.5395 \\ (0.0283) & (0.1204) \end{array}\right) { diag }\left(\left(\begin{array}{c} 0.6475 \\ (0.0383) \\ 0.2658 \\ (0.0721) \end{array}\right)^{\prime}\right)
    Thursday \left(\begin{array}{c} 1.4214 \\ (0.5815) \\ 0.0053 \\ (0.0081) \end{array}\right) \left(\begin{array}{ll} 0.0674 & 0.3020 \\ (0.0298) & (0.0526) \\ 0.0708 & 0.4969 \\ (0.0156) & (0.1282) \end{array}\right) {diag}\left(\left(\begin{array}{c} 0.6959 \\ (0.0802) \\ 0.0407 \\ (0.0926) \end{array}\right)^{\prime}\right)
    Friday \left(\begin{array}{c} 0.5244 \\ (0.0099) \\ 0.2792 \\ (0.0135) \end{array}\right) \left(\begin{array}{ll} 0.0576 & 0.0664 \\ (0.0185) & (0.0551) \\ 0.0763 & 0.2715 \\ (0.0074) & (0.0460) \end{array}\right) {diag}\left(\left(\begin{array}{c} 0.9726 \\ (0.0033) \\ 0.2879 \\ (0.0496) \end{array}\right)^{\prime}\right)

     | Show Table
    DownLoad: CSV

    Several rudimentary descriptive statistics are provided for the two \log-return series in Tables 3 and 4.

    In Tables 5 and 6, we unveil the outcomes from the QML parameter estimation conducted for the periodic CCC-AGARCH(1, 1) vis-à-vis periodic CCC-GARCH(1, 1) models. The results distinctly showcase that the periodically estimated models exhibit periodic stationarity. Notably, the measurement invariance estimates of the periodic CCC - AGARCH(1, 1) model are significantly smaller than those obtained from the periodic CCC-GARCH(1, 1) model fitted by Bibi (2018, [5]). Furthermore, the empirical coverages of the prediction intervals based on the periodic CCC-AGARCH(1, 1) model tend to closely align with the nominal coverages, unlike those derived from the periodic (EUR/DZD, USD/DZD) -CCC model. These observations suggest that the periodic GARCH(1, 1)-5 model, fitted to daily CCC log returns time series, evinces enhanced precision and superior forecasting efficacy relative to its AGARCH(1, 1)- \widehat{\underline{a}}_{0} counterpart. This comparative analysis underscores the significant contribution of Patton (2011, [29]) to developing volatility forecasting models, offering improved efficiency and accuracy.

    This research marks a significant advancement in asymmetric multivariate stochastic difference volatility modeling. Through meticulous analysis, we have successfully extended and enhanced various probabilistic and statistical properties that have been a focal point in the prior literature. A key achievement of this study is the establishment of periodic stationarity and periodic ergodicity within the model, underscoring its dynamic and adaptable nature. Furthermore, our in-depth exploration of the Quasi-Maximum Likelihood Estimator (QMLE) reveals its robust consistency and asymptotic normality. This rigorous examination not only fortifies the theoretical foundation of QMLE but also underscores its reliability and efficiency in practical scenarios. The analytical insights gained here significantly bolster the estimator's credibility in complex stochastic modeling contexts.

    The practical implications of our findings are illustrated through a series of applications, demonstrating the model's versatility and relevance in real-world situations. These applications validate our theoretical results and showcase the model's potential in addressing a range of challenging problems in stochastic volatility modeling.

    In summary, this study not only enriches the theoretical framework of stochastic volatility modeling, but also paves the way for future investigations. By pushing the boundaries of current understanding and application, it opens new avenues for research and development in this ever-evolving domain. The implications of this work extend beyond the immediate scope of this study, offering valuable insights and tools for both researchers and practitioners in the field of financial econometrics and beyond.

    Omar Alzeley: Methodology, Validation, Formal analysis, Investigation, Visualization, Writing-review and Editing; Ahmed Ghezal: Software, Validation, Resources, Data curation, Writing-original draft preparation, Supervision, Project administration. All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    We would like to express our gratitude to the Editor-in-Chief of the journal, the Assistant Editors, and the three anonymous referees for their constructive comments, valuable suggestions, and remarks, which were instrumental in improving the final version of the paper. Additionally, we extend our thanks to Prof. Dr. Amirouche Bouchelaghem, Director of the University Center, and our colleague, Dr. Khalil Zerrari, for their important encouragement and support.

    The authors declare no conflicts of interest.

    Proof of Theorem 1. The proof of the sufficient condition is analogous to that provided by Bougerol and Picard [10,11], since E\left\{ \widetilde {\underline{e}}_{t}^{\otimes2}\right\} < \infty , where

    \widetilde {\underline{e}}_{t}: = (\left( \widetilde{\underline{\eta}}_{st}^{+2}\right) ^{\prime},\dots, \left( \widetilde{\underline{\eta}}_{st+s-1}^{+2}\right) ^{\prime},\left( \widetilde{\underline{\eta}}_{st}^{-2}\right) ^{\prime },\dots, \left( \widetilde{\underline{\eta}}_{st+s-1}^{-2}\right) ^{\prime })^{\prime},

    hence E\left\{ \log^{+}\left\Vert \Omega_{t}\right\Vert \right\} < \infty. By employing Cauchy's root test when \gamma_{\Omega} < 0 , it follows that the series \underline{\Gamma}_{t}\left(2\right) converges almost surely for all t . Consequently, an SPS solution to the model (2.1) is attained, denoted as \underline {X}_{t} = \left\{ diag\left(\underline{\Lambda}_{2q+1, t}\right) \right\} ^{1/2}M^{1/2}\underline{\eta}_{t} , where \underline{\Lambda}_{2q+1, t} is the subvector of \underline{\Lambda}_{t} corresponding to the \left(2q+1\right) ^{th} element. This solution is thus nonanticipative and PE . To prove the uniqueness, let (\widetilde{\underline{\Gamma}}_{t}) be another non-negative and SS solution of (2.4). For all positive integers n\geq0 , \underline{\Gamma}_{t} = \left\{ \prod \limits_{j = 0}^{n}\Omega_{t-j}\right\} \underline{\Gamma}_{t-n-1} +\widetilde{\underline{\Gamma}}_{t}\left(n\right) , where \widetilde {\underline{\Gamma}}_{t}\left(n\right) : = \underline{\Xi}_{t}+\sum \limits_{i = 0}^{n}\left\{ \prod\limits_{j = 0}^{i}\Omega_{t-j}\right\} \underline{\Xi}_{t-i-1}. Then \left\Vert \underline{\Gamma}_{t} -\widetilde{\underline{\Gamma}}_{t}\right\Vert \leq\left\Vert \prod \limits_{j = 0}^{n}\Omega_{t-j}\right\Vert \left\Vert \underline{\Gamma} _{t-n-1}\right\Vert +\left\Vert \widetilde{\underline{\Gamma}}_{t}\left(n\right) -\widetilde{\underline{\Gamma}}_{t}\right\Vert. The second term on the right-hand side converges to 0 almost surely as n approaches infinity. Moreover, we have \left\Vert \prod\limits_{j = 0}^{n}\Omega_{t-j}\right\Vert \longrightarrow0 with a probability of 1 when n\longrightarrow\infty and \left\Vert \prod\limits_{j = 0}^{n}\Omega_{t-j}\right\Vert \left\Vert \underline{\Gamma}_{t-n-1}\right\Vert \longrightarrow0 in probability as n\longrightarrow\infty. We have shown that \underline{\Gamma} _{t} = \widetilde{\underline{\Gamma}}_{t} for any t , a.s. Now, we give a simple direct proof. Since the necessary part is easy, see Bougerol and Picard [10], we prove only \left\Vert \prod\limits_{j = 0}^{n}\Delta_{t-j}\right\Vert \longrightarrow0 as n\longrightarrow\infty . For this, it will suffice to show that

    \left\{ \prod\limits_{j = 0}^{n}\Delta_{t-j}\right\} \varphi_{k} \longrightarrow O\text{ a.s. when }n\longrightarrow\infty, (A.1)

    for all k = 1, \dots, 2q+p , where \varphi_{k} = \underline{\psi}_{k}\otimes I_{\left(r\right) } and \underline{\psi}_{k} represents the k^{th} element of the canonical basis of \mathbb{R}^{2q+p}. Indeed, assume that model (2.3) admits a causal, SPS solution \left(\underline{\Lambda}_{t}\right) . We exploit the non-negativity of the coefficients of \Delta_{t} and \underline{\Lambda}_{t} , from which it follows that for all n\geq2, \underline{\Lambda}_{t}\geq\underline{\Pi} _{t}+\sum\limits_{k = 0}^{n}\left\{ \prod\limits_{j = 0}^{k}\Delta_{t-j}\right\} \underline{\Pi}_{t-k-1} \forall n implies that \left\{ \prod \limits_{j = 0}^{n}\Delta_{t-j}\right\} \underline{\Pi}_{t-n-1}\longrightarrow \underline{O} a.s. as n\longrightarrow\infty. Using the relationship

    \underline{\Pi}_{t-n-1} = \varphi_{1}G_{t-n-1}^{+}\underline {a}_{0}\left( t-n-1\right) +\varphi_{q+1}G_{t-n-1}^{-}\underline{a} _{0}\left( t-n-1\right) +\varphi_{2q+1}\underline{a}_{0}\left( t-n-1\right),

    we have

    \left\{ \prod\limits_{j = 0}^{n}\Delta_{t-j}\right\} \varphi_{1}G_{t-n-1}^{+}\underline{a}_{0}\left( t-n-1\right) \longrightarrow \underline{O}, \left\{ \prod\limits_{j = 0}^{n}\Delta_{t-j}\right\} \varphi_{q+1}G_{t-n-1}^{-}\underline{a}_{0}\left( t-n-1\right) \longrightarrow\underline{O}

    and

    \left\{ \prod\limits_{j = 0}^{n} \Delta_{t-j}\right\} \varphi_{2q+1}\underline{a}_{0}\left( t-n-1\right) \longrightarrow\underline{O}\; a.s.\; as\; n\longrightarrow\infty.

    Since the components of \underline{a}_{0}\left(t-n-1\right) are strictly positive, (A.1) thus holds for n = 2q+1 . Using

    \Delta _{t-n}\varphi_{2q+k} = G_{t-n}^{+}C_{k}\left( t-n\right) \varphi_{1} +G_{t-n}^{-}C_{k}\left( t-n\right) \varphi_{q+1}+C_{k}\left( t-n\right) \varphi_{2q+1}+\varphi_{2q+k+1}, k = 1,...,p

    with the convention that

    \varphi_{2q+p+1} = O, for k = 1 we obtain O = \lim\limits_{n\longrightarrow\infty }\left\{ \prod\limits_{j = 0}^{n}\Delta_{t-j}\right\} \varphi_{2q+1}\geq \lim\limits_{n\longrightarrow\infty}\left\{ \prod\limits_{j = 0}^{n-1}\Delta _{t-j}\right\} \varphi_{2q+2}\geq O.

    In this context, the inequalities are considered componentwise. Moreover, (A.1) is valid for k = q+2 , and, by induction, for k = 2q+l, l = 1, \dots, p . Moreover, since

    \begin{align*} \Delta_{t-n}\varphi_{2q}& = G_{t-n}^{+}B_{q}\left( t-n\right) \varphi _{1}+G_{t-n}^{-}B_{q}\left( t-n\right) \varphi_{q+1}+B_{q}\left( t-n\right) \varphi_{2q+1} (resp. \Delta_{t-n}\varphi_{q}\\ & = G_{t-n}^{+} A_{q}\left( t-n\right) \varphi_{1}+G_{t-n}^{-}A_{q}\left( t-n\right) \varphi_{q+1}+A_{q}\left( t-n\right) \varphi_{2q+1}), \end{align*}

    (A.1) holds for k = 2q (resp. k = q ), we reach a similar finding for the remaining values of k through an upward recursion.

    Proof of Corollary 1. Since \Delta_{t} is non-negative, given that all its elements are non-negative, it is evident that \gamma_{\Omega} corresponding to the nonrandom sequence of matrices \left(\Omega_{t}^{0}\right) attained through substitution of the matrices A_{j}\left(t\right) and B_{j}\left(t\right) with a null matrix in \Delta_{t}. In a different sense, it implies \log\rho\left(\prod\limits_{v = 0}^{s-1}C_{s-v}\right) \leq \gamma_{\Omega} < 0. Consequently, the conclusion stated in the corollary is affirmed.

    Proof of Proposition 1. To prove (a). By Kingman's subadditive ergodic theorem, we get

    \lim\limits_{t\longrightarrow\infty}\frac{1}{t}\log\left\Vert \prod\limits_{k = 0} ^{t-1}\Omega_{t-k}\right\Vert \leq\log\left( \rho\left( E\left\{ \Omega _{1}\right\} \right) \right),

    which completes the proof of (a). We now prove (b). By Jensen's inequality, we get

    \gamma_{\Omega}: = \inf\limits_{t\geq1}\left\{ E\left\{ \log\left\Vert \prod\limits_{i = 0}^{t-1}\Omega_{t-i}\right\Vert ^{1/t}\right\} \right\} \leq E\left\{ \log\left\Vert \prod\limits_{i = 0}^{t-1}\Omega_{t-i}\right\Vert \right\} \leq\log E\left\{ \left\Vert \prod\limits_{i = 0}^{t-1}\Omega _{t-i}\right\Vert \right\} a.s.

    Furthermore, the rest is analogous to that of Lemma 2.3 of Berkes et al.[4] in the univariate case, and [5] in the multivariate case that the SPS solution conforms to E\left\{ \left\Vert \underline {\Lambda}_{t}\right\Vert ^{\varsigma}\right\} < \infty for some \varsigma > 0 . The result is a natural consequence of \left\Vert \underline{X} _{t}\right\Vert \leq\left\Vert \underline{\Lambda}_{t}\right\Vert and \left\Vert \underline{\sigma}_{t}\right\Vert \leq\left\Vert \underline {\Lambda}_{t}\right\Vert .

    Proof of Theorem 2. The arguments in the proof closely resemble those found in [5].

    Proof of Lemma 1. Rewrite (2.2) in a vector form as

    \underline{\Sigma}_{t}\left( \underline{\theta}\right) = C_{t}\underline {\Sigma}_{t-1}\left( \underline{\theta}\right) +\underline{\pi} _{t}, (A.2)

    where

    \underline{\Sigma}_{t}^{\prime}\left( \underline{\theta}\right) = \left( \underline{\sigma}_{t}^{\prime}\left( \underline{\theta}\right) ,...,\underline{\sigma}_{t-p+1}^{\prime}\left( \underline{\theta}\right) \right)

    and

    \underline{\pi}_{t}^{\prime} = \left( \left( \underline{a} _{0}\left( t\right) +\sum\limits_{i = 1}^{q}\left( A_{i}\left( t\right) \underline{X}_{t-i}^{+}+B_{i}\left( t\right) \underline{X}_{t-i}^{-}\right) \right) ^{\prime},\underline{O}_{\left( r\right) }^{\prime},...,\underline {O}_{\left( r\right) }^{\prime}\right).

    Proof of \mathbf{(I)} : Initial values are asymptotically irrelevant. In view of the last part imposed in assumption A2 and Corollary 1, we have

    \sup\limits_{\underline{\theta}\in\Theta}\rho\left( \prod\limits_{v = 0}^{s-1} C_{s-v}\right) < 1. (A.3)

    Iteratively using (A.2), as in the univariate case, we have

    \underline{\Sigma}_{t}\left( \underline{\theta}\right) = \left\{ \prod\limits_{j = 0}^{t}C_{t-j}\right\} \underline{\Sigma}_{0}\left( \underline{\theta}\right) +\sum\limits_{i = 0}^{t-1}\left\{ \prod \limits_{j = 0}^{i-1}C_{t-j}\right\} \underline{\pi}_{t-i}, t\in \mathbb{Z}.

    If we represent the vectors attained from certain initial values as \widetilde{\underline{\Sigma}}_{t}\left(\underline{\theta}\right) and \widetilde{\underline{\pi}}_{t} , then we obtain

    \widetilde{\underline {\Sigma}}_{t}\left( \underline{\theta}\right) = \left\{ \prod\limits_{j = 0} ^{t}C_{t-j}\right\} \widetilde{\underline{\Sigma}}_{0}\left( \underline {\theta}\right) +\sum\limits_{i = 0}^{t-q-1}\left\{ \prod\limits_{j = 0} ^{i-1}C_{t-j}\right\} \underline{\pi}_{t-i}+\sum\limits_{i = t-q}^{t-1}\left\{ \prod\limits_{j = 0}^{i-1}C_{t-j}\right\} \widetilde{\underline{\pi}}_{t-i}, t\in \mathbb{Z}.

    Hence, we deduce that a.s.

    \sup\limits_{\underline{\theta}\in\Theta}\left\Vert \widetilde{\underline{\Sigma} }_{t}\left( \underline{\theta}\right) -\underline{\Sigma}_{t}\left( \underline{\theta}\right) \right\Vert \leq\lambda\omega^{t},\forall t,

    where \lambda > 0 and \omega\in\left(0, 1\right) , which implies that \sup_{\underline{\theta}\in\Theta}\left\Vert \widetilde{V}_{t}\left(\underline{\theta}\right) -V\left(\underline{\theta}\right) \right\Vert \leq\lambda\omega^{t} \forall t . Since H_{t}^{-1}\left(\underline {\theta}\right) \leq\left\{ \min_{j}\sigma_{t, j}^{2}\right\} ^{-1} , we have

    \sup\limits_{\underline{\theta}\in\Theta}\left\Vert \widetilde{V}_{t}^{-1}\left( \underline{\theta}\right) \right\Vert \leq\sup\limits_{\underline{\theta}\in\Theta }\left\Vert \widetilde{H}_{t}^{-1}\left( \underline{\theta}\right) \right\Vert ^{2}\left\Vert M^{-1}\right\Vert \leq\sup\limits_{\underline{\theta} \in\Theta}\left\{ \min\limits_{j}a_{0,j}\left( t\right) \right\} ^{-2}\left\Vert M^{-1}\right\Vert \leq\lambda,

    Similarly, we obtain \sup_{\underline{\theta}\in\Theta}\left\Vert V_{t} ^{-1}\left(\underline{\theta}\right) \right\Vert \leq\lambda .

    Now,

    \sup\limits_{\underline{\theta}\in\Theta}\left\vert \left( \widetilde{L}_{Ns} -L_{Ns}\right) \left( \underline{\theta}\right) \right\vert \leq\frac{1} {N}\sum\limits_{t = 1}^{Ns}\sup\limits_{\underline{\theta}\in\Theta}\left\vert \underline{X}_{t}^{\prime}\left( \widetilde{V}_{t}^{-1}-V_{t}^{-1}\right) \left( \underline{\theta}\right) \underline{X}_{t}\right\vert +\frac{1} {N}\sum\limits_{t = 1}^{Ns}\sup\limits_{\underline{\theta}\in\Theta}\left\vert \log\left( \left. \left\vert \widetilde{V}_{t}\left( \underline{\theta }\right) \right\vert \right/ \left\vert V_{t}\left( \underline{\theta }\right) \right\vert \right) \right\vert .

    The first sum can be written as

    \begin{align*} \frac{1}{N}\sum\limits_{t = 1}^{Ns}\sup\limits_{\underline{\theta}\in\Theta}\left\vert \underline{X}_{t}^{\prime}\left( \widetilde{V}_{t}^{-1}\left( \widetilde {V}_{t}-V_{t}\right) V_{t}^{-1}\right) \left( \underline{\theta}\right) \underline{X}_{t}\right\vert & = \frac{1}{N}\sum\limits_{t = 1}^{Ns} \sup\limits_{\underline{\theta}\in\Theta}\left\vert Tr\left( \left( \widetilde {V}_{t}^{-1}\left( \widetilde{V}_{t}-V_{t}\right) V_{t}^{-1}\right) \left( \underline{\theta}\right) \underline{X}_{t}\underline{X}_{t}^{\prime}\right) \right\vert \\ & \leq\frac{\lambda}{N}\sum\limits_{t = 1}^{Ns}\sup\limits_{\underline{\theta} \in\Theta}\left\Vert \widetilde{V}_{t}^{-1}\left( \underline{\theta}\right) \right\Vert \left\Vert \left( \widetilde{V}_{t}-V_{t}\right) \left( \underline{\theta}\right) \right\Vert \left\Vert V_{t}^{-1}\left( \underline{\theta}\right) \right\Vert \left\Vert \underline{X}_{t} \underline{X}_{t}^{\prime}\right\Vert \\ & \leq\frac{\lambda}{N}\sum\limits_{t = 1}^{Ns}\omega^{t}\left\Vert \underline{X}_{t}\underline{X}_{t}^{\prime}\right\Vert \longrightarrow0\text{ as }N\longrightarrow\infty, \end{align*}

    Using the Cesaro lemma, the Borel–Cantelli lemma, the Markov inequality, and by applying Proposition 1, we have

    \sum\limits_{t\geq1}P\left( \omega^{t}\underline{X}_{t}\underline{X}_{t}^{\prime} > \varepsilon\right) \leq\sum\limits_{t\geq1}\dfrac{\omega^{t\tau}E\left\{ \left\Vert \underline{X}_{t}\right\Vert ^{\tau}\right\} }{\varepsilon^{\tau}} < \infty,

    which implies \omega^{t}\underline{X}_{t}^{\prime}\underline{X}_{t} \overset{a.s.}{\longrightarrow}0 . Now, using the inequality for y+1\geq0 , y\geq\log\left(1+y\right) , we have

    \log\left( \left. \left\vert V_{t}\left( \underline{\theta}\right) \right\vert \right/ \left\vert \widetilde{V}_{t}\left( \underline{\theta }\right) \right\vert \right) = \log\left\vert I_{\left( r\right) }+\left( \left( V_{t}-\widetilde{V}_{t}\right) \widetilde{V}_{t}^{-1}\right) \left( \underline{\theta}\right) \right\vert \leq r\left\Vert \left( V_{t} -\widetilde{V}_{t}\right) \left( \underline{\theta}\right) \right\Vert \left\Vert \widetilde{V}_{t}^{-1}\left( \underline{\theta}\right) \right\Vert ,

    and, by symmetry,

    \log\left\vert \widetilde{V}_{t}\left( \underline{\theta }\right) \right\vert -\log\left\vert V_{t}\left( \underline{\theta}\right) \right\vert \leq r\left\Vert V_{t}\left( \underline{\theta}\right) -\widetilde{V}_{t}\left( \underline{\theta}\right) \right\Vert \left\Vert V_{t}^{-1}\left( \underline{\theta}\right) \right\Vert ,

    so we deduce that the second sum tends to 0 . Therefore, we have demonstrated that

    \sup\limits_{\underline{\theta}\in\Theta}\left\vert \left( \widetilde{L} _{Ns}-L_{Ns}\right) \left( \underline{\theta}\right) \right\vert \underset{N\longrightarrow\infty}{\overset{a.s.}{\longrightarrow}}0.

    Proof of \mathbf{(II)} : Parameter identifiability.. Suppose that, for some \underline{\theta}\neq\underline{\theta}_{0}, \underline{\sigma }_{t}\left(\underline{\theta}\right) = \underline{\sigma}_{t}\left(\underline{\theta}_{0}\right) a.s. and M = M_{0}.

    Consequently, it immediately follows that \underline{m} = \underline{m}_{0} , and by leveraging the invertibility of the polynomial \left(\mathcal{C}_{0, v}\left(z\right) \right) _{1\leq v\leq s} under the assumption A2, then we have a.s. \forall v = 1, ..., s,

    \begin{align*} \mathcal{C}_{0,v}^{-1}\left( L\right) \underline{a}_{0,0}\left( v\right) -\mathcal{C}_{v}^{-1}\left( L\right) \underline{a}_{0}\left( v\right) & = \left( \mathcal{C}_{v}^{-1}\left( L\right) \mathcal{A}_{v}\left( L\right) -\mathcal{C}_{0,v}^{-1}\left( L\right) \mathcal{A}_{0,v}\left( L\right) \right) \underline{X}_{ns+v}^{+}\\ & +\left( \mathcal{C}_{v}^{-1}\left( L\right) \mathcal{B}_{v}\left( L\right) -\mathcal{C}_{0,v}^{-1}\left( L\right) \mathcal{B}_{0,v}\left( L\right) \right) \underline{X}_{ns+v}^{-}\\ & : = \mathcal{P}_{v}\left( L\right) \underline{X}_{ns+v}^{+}+\mathcal{Q} _{v}\left( L\right) \underline{X}_{ns+v}^{-}, \end{align*}

    where L denotes the lag operator,

    \mathcal{P}_{v}\left( L\right) = \sum\nolimits_{k\geq1}P_{k}\left( v\right) L^{k}, \mathcal{Q}_{v}\left( L\right) = \sum\nolimits_{k\geq1}Q_{k}\left( v\right) L^{k}

    with

    \mathcal{P}_{v}\left( 0\right) = P_{0}\left( v\right) = \mathcal{Q} _{v}\left( 0\right) = Q_{0}\left( v\right) = O_{\left( r\right) },

    and isolating the terms that are functions of \underline{\eta}_{t-1} ,

    P_{1}\left( t\right) G_{t-1}^{+}\underline{\sigma}_{t-1}+Q_{1}\left( t\right) G_{t-1}^{-}\underline{\sigma}_{t-1}\in\sigma\left( \underline{\eta }_{t-2},\underline{\eta}_{t-3},...\right).

    Since P_{1}\left(t\right) G_{t-1}^{+}+Q_{1}\left(t\right) G_{t-1}^{-} is independent of this \sigma- field and \underline{\sigma}_{t-1} and \underline{\sigma}_{t-1} > 0 , we have

    P_{1}\left( t\right) G_{t-1}^{+}+Q_{1}\left( t\right) G_{t-1}^{-} = T

    for some constant matrix T , the matrix equality for the element \left(k, l\right) writes \left(P_{1}\left(t\right) \right) \left(k, l\right) \widetilde{\underline{\eta}}_{t-1}^{+2}\left(l\right) +\left(Q_{1}\left(t\right) \right) \left(k, l\right) \widetilde {\underline{\eta}}_{t-1}^{-2}\left(l\right) = T\left(k, l\right). If \left(P_{1}\left(t\right) \right) \left(k, l\right) \left(Q_{1}\left(t\right) \right) \left(k, l\right) \neq0. The contradiction with Assumption A3 arises from the fact that \widetilde {\underline{\eta}}_{t-1}\left(l\right) takes at most two distinct values.

    If \left(P_{1}\left(t\right) \right) \left(k, l\right) \neq0 and \left(Q_{1}\left(t\right) \right) \left(k, l\right) = 0, thus \left(P_{1}\left(t\right) \right) \left(k, l\right) \widetilde {\underline{\eta}}_{t-1}^{+2}\left(l\right) = T\left(k, l\right), which entails T\left(k, l\right) = 0, because P\left(\widetilde{\underline {\eta}}_{t-1}^{-}\left(l\right) > 0\right) \neq0, and thus \widetilde {\underline{\eta}}_{t-1}^{-}\left(l\right) \overset{a.s.}{ = }0 , which contradicts A3. Therefore, we have P_{1}\left(t\right) = Q_{1}\left(t\right) = 0 . Similarly in this way, we show that \mathcal{P}_{v}\left(L\right) = \mathcal{Q}_{v}\left(L\right) = 0 for all v = 1, \dots, s . Therefore, in view of A4, we have \underline{a} _{0}\left(v\right) = \underline{a}_{0, 0}\left(v\right), \underline {\alpha}\left(v\right) = \underline{\alpha}_{0}\left(v\right), \underline{\beta}\left(v\right) = \underline{\beta}_{0}\left(v\right) and \underline{\gamma}\left(v\right) = \underline{\gamma}_{0}\left(v\right) for all v = 1, ..., s . Hence \underline{\theta} = \underline{\theta }_{0} . Thus, we have successfully established \mathbf{(II)} .

    Proof of \mathbf{(III)} : The limit criterion attains its minimum at \underline{\theta}_{0} .. At \underline{\theta}_{0} , Jensen's inequality and Proposition 1 entail that

    \begin{align*} \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \log\left\vert V_{st+v}\left( \underline{\theta}_{0}\right) \right\vert \right\} & = \frac{r}{\tau}\sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \log\left\vert V_{st+v}\left( \underline{\theta}_{0}\right) \right\vert ^{\frac{\tau}{r}}\right\} \\ &\leq\frac{r}{\tau}\sum\nolimits_{v = 1}^{s}\log E_{\underline{\theta}_{0}}\left\{ \left\Vert V_{st+v}\left( \underline {\theta}_{0}\right) \right\Vert ^{\tau}\right\} \\ & \leq\frac{r}{\tau}\sum\nolimits_{v = 1}^{s}\log E_{\underline{\theta}_{0} }\left\{ \left\Vert M\right\Vert ^{\tau}\left\Vert H_{st+v}\left( \underline{\theta}_{0}\right) \right\Vert ^{2\tau}\right\} \\ &\leq\lambda +\frac{r}{\tau}\sum\nolimits_{v = 1}^{s}\log E_{\underline{\theta}_{0}}\left\{ \left\Vert H_{st+v}\left( \underline{\theta}_{0}\right) \right\Vert ^{2\tau }\right\} \\ & \leq\lambda+\frac{r}{\tau}\sum\nolimits_{v = 1}^{s}\log E_{\underline{\theta }_{0}}\left\{ \max\limits_{i}\left\{ \sigma_{i\text{ }st+v}\left( \underline {\theta}_{0}\right) \right\} ^{\tau}\right\} \\ &\leq\lambda+\frac{r}{\tau} \sum\nolimits_{v = 1}^{s}\log E_{\underline{\theta}_{0}}\left\{ \left\Vert \underline{\sigma}_{st+v}\left( \underline{\theta}_{0}\right) \right\Vert ^{\tau}\right\} \\ & < \infty. \end{align*}

    As a result,

    \begin{align*} \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ l_{st+v}\left( \underline{\theta}_{0}\right) \right\} & = \sum\nolimits_{v = 1} ^{s}E_{\underline{\theta}_{0}}\left\{ \underline{\eta}_{st+v}^{\prime }V_{st+v}^{1/2\prime}\left( \underline{\theta}_{0}\right) V_{st+v} ^{-1}\left( \underline{\theta}_{0}\right) V_{st+v}^{1/2}\left( \underline{\theta}_{0}\right) \underline{\eta}_{st+v}+\log\left\vert V_{st+v}\left( \underline{\theta}_{0}\right) \right\vert \right\} \\ & = r+\sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \log\left\vert V_{st+v}\left( \underline{\theta}_{0}\right) \right\vert \right\} < \infty. \end{align*}

    We now have

    \begin{align*} &\sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ l_{st+v}\left( \underline{\theta}\right) -l_{st+v}\left( \underline{\theta}_{0}\right) \right\} = \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \log\frac{\left\vert V_{st+v}\left( \underline{\theta}\right) \right\vert }{\left\vert V_{st+v}\left( \underline{\theta}_{0}\right) \right\vert }\right\} \\ & +\sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \underline{\eta }_{st+v}^{\prime}\left( V_{st+v}^{1/2}\left( \underline{\theta}_{0}\right) V_{st+v}^{-1}\left( \underline{\theta}\right) V_{st+v}^{1/2}\left( \underline{\theta}_{0}\right) -I_{\left( r\right) }\right) \underline {\eta}_{st+v}\right\} \\ & = \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \log\left\vert V_{st+v}\left( \underline{\theta}\right) \right\vert \left\vert V_{st+v}^{-1}\left( \underline{\theta}_{0}\right) \right\vert \right\} \\ & +\sum\nolimits_{v = 1}^{s}Tr\left( E_{\underline{\theta}_{0}}\left\{ \left( V_{st+v}^{1/2}\left( \underline{\theta}_{0}\right) V_{st+v} ^{-1}\left( \underline{\theta}\right) V_{st+v}^{1/2}\left( \underline {\theta}_{0}\right) -I_{\left( r\right) }\right) E\left\{ \underline {\eta}_{st+v}\underline{\eta}_{st+v}^{\prime}\right\} \right\} \right) \\ & = \sum\nolimits_{v = 1}^{s}\left( E_{\underline{\theta}_{0}}\left\{ \log\left\vert V_{st+v}\left( \underline{\theta}\right) \right\vert \left\vert V_{st+v}^{-1}\left( \underline{\theta}_{0}\right) \right\vert \right\} +E_{\underline{\theta}_{0}}\left\{ Tr\left( V_{st+v}\left( \underline{\theta}_{0}\right) V_{st+v}^{-1}\left( \underline{\theta}\right) -I_{\left( r\right) }\right) \right\} \right) \\ & = \sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \sum \nolimits_{j = 1}^{r}\left( \delta_{j}\left( v\right) -1-\log\delta _{j}\left( v\right) \right) \right\} , \end{align*}

    where \delta_{j}\left(v\right) is the positive eigenvalues of V_{st+v}\left(\underline{\theta}\right) V_{st+v}^{-1}\left(\underline{\theta}_{0}\right), because y-1\geq\log y, \forall y > 0, where equality holds iff y = 1 , thus we obtain \sum\nolimits_{v = 1} ^{s}E_{\underline{\theta}_{0}}\left\{ l_{st+v}\left(\underline{\theta }\right) \right\} -\sum\nolimits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ l_{st+v}\left(\underline{\theta}_{0}\right) \right\} \geq0, and hence the inequality is strict unless if, for all j and v , \delta_{j}\left(v\right) \overset{a.s.}{ = }1 , meaning that V_{st+v}\left(\underline {\theta}\right) \overset{a.s.}{ = }V_{st+v}\left(\underline{\theta} _{0}\right). Therefore, we have \underline{\sigma}_{st+v}\left(\underline{\theta}\right) \overset{a.s.}{ = }\underline{\sigma}_{st+v}\left(\underline{\theta}_{0}\right) , v = 1, ..., s and M = M_{0}\ and thus \underline{\theta} = \underline{\theta}_{0}.

    Proof. To demonstrate \mathbf{(IV), } \forall\underline{\theta}\in\Theta and every integer k , let \mathcal{V}_{k}\left(\underline{\theta}\right) be an open sphere of centre \underline{\theta} and radius \dfrac{1}{k} . Utilizing \left(\mathbf{I}\right), we obtain

    \begin{align*} \underset{N\rightarrow \infty}{\underline{\lim}}\underset{\underline{\theta }^{\ast}\in\Theta\cap\mathcal{V}_{k}\left( \underline{\theta}\right) }{\inf }\left( \widetilde{L}_{Ns}\left( \underline{\theta}^{\ast}\right) \right) & \geq\underset{N\rightarrow \infty}{\underline{\lim}}\underset{\underline {\theta}^{\ast}\in\Theta\cap\mathcal{V}_{k}\left( \underline{\theta}\right) }{\inf}\left( L_{Ns}\left( \underline{\theta}^{\ast}\right) \right) -\underset{N\rightarrow \infty}{\overline{\lim}}\underset{\underline{\theta} \in\Theta}{\sup}\left\vert \left( L_{Ns}-\widetilde{L}_{Ns}\right) \left( \underline{\theta}\right) \right\vert \\ & \geq\underset{N\rightarrow \infty}{\underline{\lim}}\frac{1}{N} \sum\limits_{t = 0}^{N-1}\sum\limits_{v = 1}^{s}\underset{\underline{\theta} ^{\ast}\in\Theta\cap\mathcal{V}_{k}\left( \underline{\theta}\right) }{\inf }l_{st+v}\left( \underline{\theta}^{\ast}\right) . \end{align*}

    By applying the ergodic theorem to the i.i.d. sequence \left(\sum \limits_{v = 1}^{s}l_{st+v}\left(\underline{\theta}^{\ast}\right) \right) _{t}, where E\left\{ \sum\limits_{v = 1}^{s}l_{st+v}\left(\underline {\theta}^{\ast}\right) \right\} \in\mathbb{R}\cup\left\{ \infty\right\} (see., [8]), we have

    \underset{N\rightarrow \infty}{\lim\inf}\frac{1}{N}\sum\limits_{t = 0}^{N-1} \sum\limits_{v = 1}^{s}\underset{\underline{\theta}^{\ast}\in\Theta \cap\mathcal{V}_{k}\left( \underline{\theta}\right) }{\inf}l_{st+v}\left( \underline{\theta}^{\ast}\right) = \sum\limits_{v = 1}^{s}E_{\underline{\theta }_{0}}\left\{ \underset{\underline{\theta}^{\ast}\in\Theta\cap\mathcal{V} _{k}\left( \underline{\theta}\right) }{\inf}l_{st+v}\left( \underline {\theta}^{\ast}\right) \right\}

    and according to the Beppo-Levi theorem (see., [8]), we get

    \sum\limits_{v = 1}^{s}E_{\underline{\theta}_{0}}\left\{ \underset {\underline{\theta}^{\ast}\in\Theta\cap\mathcal{V}_{k}\left( \underline {\theta}\right) }{\inf}l_{st+v}\left( \underline{\theta}^{\ast}\right) \right\} \longrightarrow\sum\limits_{v = 1}^{s}E_{\underline{\theta}_{0} }\left\{ l_{st+v}\left( \underline{\theta}\right) \right\}

    as k\rightarrow \infty. This concludes the proof of the lemma.

    Proof of Theorem 3. The proof is now complete by an argument of compactness of \Theta and Lemma 1. Now, for all neighborhood \mathcal{V}\left(\underline{\theta}_{0}\right), we obtain

    \underset{N\rightarrow \infty}{\overline{\lim}}\underset{\underline{\theta }^{\ast}\in\mathcal{V}\left( \underline{\theta}_{0}\right) }{\inf}\left( \widetilde{L}_{Ns}\left( \underline{\theta}^{\ast}\right) \right) \leq\underset{N\rightarrow \infty}{\lim}\widetilde{L}_{Ns}\left( \underline{\theta}_{0}\right) = \underset{N\rightarrow \infty}{\lim} L_{Ns}\left( \underline{\theta}_{0}\right) = \sum\limits_{v = 1}^{s} E_{\underline{\theta}_{0}}\left\{ l_{st+v}\left( \underline{\theta} _{0}\right) \right\} .

    The compact set \Theta is reconstructed as the union of a neighborhood \mathcal{V}\left(\underline{\theta}_{0}\right) and a set of neighborhoods \mathcal{V}\left(\underline{\theta}\right) , where \underline{\theta}\in\Theta\diagdown\mathcal{V}\left(\underline{\theta} _{0}\right) . Consequently, there exists a finite sub-covering of \Theta by \mathcal{V}\left(\underline{\theta}_{0}\right), \mathcal{V}\left(\underline{\theta}_{1}\right), ..., \mathcal{V}\left(\underline{\theta} _{k}\right) such that

    \underset{\underline{\theta}^{\ast}\in\mathcal{V}\left( \underline{\theta }_{0}\right) }{\inf}\left( \widetilde{L}_{Ns}\left( \underline{\theta }^{\ast}\right) \right) = \min\limits_{j\in\left\{ 1,...,k\right\} } \underset{\underline{\theta}^{\ast}\in\Theta\cap\mathcal{V}\left( \underline{\theta}_{j}\right) }{\inf}\left( \widetilde{L}_{Ns}\left( \underline{\theta}^{\ast}\right) \right) .

    The latter relation indicates that \widehat{\underline{\theta}}_{Ns} \in\mathcal{V}\left(\underline{\theta}_{0}\right) for N is sufficiently large, thereby completing the proof of Theorem 3.

    Proof of Lemma 2. The demonstration closely follows that of [15] for the standard case, employing the SPS and PE arguments in place of the stationarity and ergodicity considerations, respectively. First and foremost, we can write the first-and second-order derivatives of l_{t}\left(\underline{\theta}\right) as follows:

    \begin{align*} \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ } {\partial\theta_{i}\left( v\right) } & = Tr\left( \underline{X} _{v}\underline{X}_{v}^{\prime}\frac{\partial H_{v}^{-1}M^{-1}H_{v}^{-1} \ }{\partial\theta_{i}\left( v\right) }\right) +2\frac{\partial \log\left\vert \det H_{v}\right\vert \ }{\partial\theta_{i}\left( v\right) }\\ & = Tr\left( \left( I_{\left( r\right) }-M^{-1}H_{v}^{-1}\underline{X} _{v}\underline{X}_{v}^{\prime}H_{v}^{-1}\right) \frac{\partial H_{v} \ }{\partial\theta_{i}\left( v\right) }H_{v}^{-1}+\left( I_{\left( r\right) }-H_{v}^{-1}\underline{X}_{v}\underline{X}_{v}^{\prime}H_{v} ^{-1}M^{-1}\right) H_{v}^{-1}\frac{\partial H_{v}\ }{\partial\theta _{i}\left( v\right) }\right) ,\\ &\quad \quad \quad 1\leq i\leq\widetilde{r},\\ \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ } {\partial\theta_{i}} & = Tr\left( \left( I_{\left( r\right) }-M^{-1} H_{v}^{-1}\underline{X}_{v}\underline{X}_{v}^{\prime}H_{v}^{-1}\right) M^{-1}\frac{\partial M\ }{\partial\theta_{i}}\right) ,\widetilde{r}+1\leq i\leq l,\\ \frac{\partial^{2}l_{st+v}\left( \underline{\theta}_{0}\right) \ } {\partial\theta_{i}\left( v\right) \partial\theta_{j}\left( v\right) } & = \sum\limits_{k = 1}^{3}Tr\left( L_{k}\left( v\right) \right) ,1\leq i,j\leq\widetilde{r},\frac{\partial^{2}l_{st+v}\left( \underline{\theta} _{0}\right) \ }{\partial\theta_{i}\left( v\right) \partial\theta_{j}\left( v\right) } = \sum\limits_{k = 1}^{2}Tr\left( L_{k+3}\left( v\right) \right) ,1\leq i\leq\widetilde{r},\widetilde{r}+1\leq j\leq l,\\ \text{ }\frac{\partial^{2}l_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) \partial\theta_{j}\left( v\right) } & = Tr\left( L_{6}\left( v\right) \right) ,\widetilde{r}+1\leq i,j\leq l, \end{align*}

    for all 1\leq v\leq s , where

    \begin{align*} 2^{-1}L_{1}\left( v\right) & = H_{v}^{-1}\left( \frac{\partial^{2}H_{v} \ }{\partial\theta_{i}\left( v\right) \partial\theta_{j}\left( v\right) }-\frac{\partial H_{v}\ }{\partial\theta_{i}\left( v\right) }H_{v}^{-1} \frac{\partial H_{v}\ }{\partial\theta_{j}\left( v\right) }\right) ,\\ L_{2}\left( v\right) & = H_{v}^{-1}\left( \frac{\partial H_{v}\ } {\partial\theta_{i}\left( v\right) }H_{v}^{-1}\underline{X}_{v}\underline {X}_{v}^{\prime}H_{v}^{-1}M^{-1}+M^{-1}H_{v}^{-1}\frac{\partial H_{v} \ }{\partial\theta_{i}\left( v\right) }H_{v}^{-1}\underline{X}_{v} \underline{X}_{v}^{\prime}\right) H_{v}^{-1}\frac{\partial H_{v}\ } {\partial\theta_{j}\left( v\right) }\\ & -2^{-1}H_{v}^{-1}\underline{X}_{v}\underline{X}_{v}^{\prime}H_{v} ^{-1}M^{-1}L_{1}\left( v\right) ,\\ L_{4}\left( v\right) & = M^{-1}H_{v}^{-1}\frac{\partial H_{v}\ } {\partial\theta_{i}\left( v\right) }H_{v}^{-1}\underline{X}_{v}\underline {X}_{v}^{\prime}H_{v}^{-1}M^{-1}\frac{\partial M\ }{\partial\theta_{j}},\\ L_{6}\left( v\right) & = M^{-1}\frac{\partial M\ }{\partial\theta_{i} }M^{-1}H_{v}^{-1}\underline{X}_{v}\underline{X}_{v}^{\prime}H_{v}^{-1} M^{-1}\frac{\partial M\ }{\partial\theta_{j}}+M^{-1}H_{v}^{-1}\underline {X}_{v}\underline{X}_{v}^{\prime}H_{v}^{-1}M^{-1}\left( \frac{\partial M\ }{\partial\theta_{i}}M^{-1}\frac{\partial M\ }{\partial\theta_{j}} -\frac{\partial^{2}M\ }{\partial\theta_{i}\partial\theta_{j}}\right) \\ & -M^{-1}\frac{\partial^{2}M\ }{\partial\theta_{i}\partial\theta_{j}} -M^{-1}\frac{\partial M\ }{\partial\theta_{i}}M^{-1}\frac{\partial M\ }{\partial\theta_{j}}, \end{align*}

    where L_{3}\left(v\right) (resp. L_{5}\left(v\right) ) is obtained by permuting \underline{X}_{v}\underline{X}_{v}^{\prime} and M^{-1} (resp. \dfrac{\partial H_{v}\ }{\partial\theta_{i}\left(v\right) } ) in L_{4}\left(v\right) . To prove \mathbf{(a)} it would be appropriate to use the Cauchy–Schwarz inequality, from which we obtain

    \begin{align*} E_{\underline{\theta}_{0}}\left\{ \left\Vert \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) } \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ } {\partial\theta_{j}\left( v\right) }\right\Vert \right\} & \leq \omega\left( E_{\underline{\theta}_{0}}\left\{ \left\Vert H_{v}^{-1}\left( \underline{\theta}_{0}\right) \frac{\partial H_{v}\ \left( \underline {\theta}_{0}\right) }{\partial\theta_{i}\left( v\right) }\right\Vert ^{2}\right\} E_{\underline{\theta}_{0}}\left\{ \left\Vert H_{v}^{-1}\left( \underline{\theta}_{0}\right) \frac{\partial H_{v}\ \left( \underline {\theta}_{0}\right) }{\partial\theta_{j}\left( v\right) }\right\Vert ^{2}\right\} \right) ^{1/2},1\leq i,j\leq\widetilde{r},\\ E_{\underline{\theta}_{0}}\left\{ \left\Vert \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) } \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ } {\partial\theta_{j}\left( v\right) }\right\Vert \right\} & \leq\omega E_{\underline{\theta}_{0}}\left\{ \left\Vert H_{v}^{-1}\left( \underline {\theta}_{0}\right) \frac{\partial H_{v}\ \left( \underline{\theta} _{0}\right) }{\partial\theta_{i}\left( v\right) }\right\Vert \right\} ,1\leq i\leq\widetilde{r},\widetilde{r}+1\leq j\leq l,\\ E_{\underline{\theta}_{0}}\left\{ \left\Vert \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) } \frac{\partial l_{st+v}\left( \underline{\theta}_{0}\right) \ } {\partial\theta_{j}\left( v\right) }\right\Vert \right\} & \leq \omega,\widetilde{r}+1\leq i,j\leq l. \end{align*}

    Now, letting \sigma_{i, t} (resp. \Sigma_{i, st+v} ) denote the i^{th} component of \underline{\sigma}_{t} (resp. \underline{\Sigma}_{st+v} ), it is thus sufficient to prove that

    E_{\underline{\theta}_{0}}\left\{ \left\vert \dfrac{1\ }{\sigma_{j,st+v}}\dfrac{\partial\sigma_{j,st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) }\right\vert ^{2}\right\} < \omega,1\leq i\leq\widetilde{r},,1\leq j\leq r.

    Then, we have a.s.

    \underset{\underline{\theta}\in\Theta}{\sup}\left\Vert \dfrac{\partial\underline{\Sigma}_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) }\right\Vert < \infty, 1\leq i\leq r, \underset{\underline{\theta}\in\Theta}{\sup}\left\Vert \dfrac {\partial\underline{\Sigma}_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) }\right\Vert < \omega\left\Vert \underline{X}_{st+v}\right\Vert ^{2}, r+1\leq i\leq\widetilde{r}-p.

    Furthermore, we have

    \begin{align*} \underline{\Sigma}_{st+v} & = \sum\limits_{j = 1}^{\infty}\left\{ \prod\limits_{k = 0}^{j-1}C_{st+v-k}\right\} \underline{\pi}_{st+v-j} +\underline{\pi}_{st+v}\\ & = \sum\limits_{j = 0}^{\infty}\left\{ \prod\limits_{k = 0}^{j}C_{s\left( t-k\right) +v}C\left( k\right) \right\} \underline{\pi}_{s\left( t-j\right) +v-m}+\underline{\pi}_{st+v},0\leq m\leq s-1, \end{align*}

    where

    C\left( m\right) = \prod\limits_{k = sm+1}^{s\left( m+1\right) -1}C_{st+v-k},0\leq m\leq j-1 and C\left( j\right) = \prod\limits_{k = sj+1} ^{sj+m-1}C_{st+v-k},0\leq m\leq s-1.

    So, we get a.s.

    \dfrac{\partial\underline{\Sigma}_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) } = \sum\limits_{j = 1}^{\infty} \sum\limits_{m = 0}^{j}\left\{ \prod\limits_{k = 0}^{m-1}C_{s\left( t-k\right) +v}C\left( k\right) \right\} C_{\left( i\right) }C\left( m\right) \left\{ \prod\limits_{k = m+1}^{j}C_{s\left( t-k\right) +v}C\left( k\right) \right\} \underline{\pi}_{s\left( t-j\right) +v-n},

    where 0\leq n\leq s-1 and C_{\left(i\right) } is a matrix whose entries are all zero except for a 1 situated in the same location as \theta _{i}\left(v\right) in C_{st+v} . Applying the inequality y^{\tau} \geq\frac{y}{1+y} for y > 0 and any \tau\in\lbrack0;1] and the fact that, almost surely \log\left\Vert \prod\limits_{k = 0}^{j-1}C_{st+v-k}\right\Vert \leq\omega+j\gamma_{\left(s\right) }\left(C\right) , we can obtain the inequalities

    \begin{align*} \theta_{i}\left( v\right) \dfrac{\partial\underline{\Sigma}_{st+v}\left( \underline{\theta}_{0}\right) \ }{\partial\theta_{i}\left( v\right) } & \leq\sum\limits_{j = 1}^{\infty}j\left\{ \prod\limits_{k = 0}^{j-1} C_{st+v-k}\right\} \underline{\pi}_{st+v-j},\\ \theta_{i}\left( v\right) \dfrac{\partial\Sigma_{n,st+v}\ }{\partial \theta_{i}\left( v\right) } & \leq\sum\limits_{j = 1}^{\infty}j\sum \limits_{m = 1}^{r}\left\{ \prod\limits_{k = 0}^{j-1}C_{st+v-k}\right\} \left( n,m\right) \underline{\pi}_{st+v-j}\left( m\right) , \end{align*}

    and, setting

    a_{0}\left( v\right) = \inf\left\{ \underline{a}_{i,0}\left( v\right) ,1\leq i\leq r\right\} , \Sigma_{i,st+v}\geq a_{0}\left( v\right) +\sum\limits_{m = 1}^{r}\left\{ \prod\limits_{k = 0}^{j-1} C_{st+v-k}\right\} \left( i,m\right) \underline{\pi}_{st+v-j}\left( m\right) \forall j,

    we obtain

    \left\vert \dfrac{\theta_{i}\left( v\right) \ }{\Sigma_{n,st+v}}\dfrac{\partial\Sigma_{n,st+v}\ }{\partial \theta_{i}\left( v\right) }\right\vert < \omega.

    Using the same technique as Francq and Zakoïan [15], we can demonstrate a result that is even more robust than the one initially stated: \forall1\leq i, j, k\leq \widetilde{r}, all 1\leq n\leq r and all m\geq0, \exists\vartheta \left(\underline{\theta}_{0}\right) of \underline{\theta}_{0} , such that

    E\left\{ \underset{\underline{\theta}\in\vartheta\left( \underline{\theta}_{0}\right) }{\sup}\left\vert \dfrac{1\ }{\sigma_{n,st+v}}\dfrac {\partial\sigma_{n,st+v}\left( \underline{\theta}\right) \ }{\partial \theta_{i}\left( v\right) }\right\vert ^{m}\right\} < \infty,
    E\left\{ \underset{\underline{\theta}\in\vartheta\left( \underline{\theta}_{0}\right) }{\sup}\left\vert \dfrac{1\ }{\sigma_{n,st+v}}\dfrac{\partial^{2} \sigma_{n,st+v}\left( \underline{\theta}\right) \ }{\partial\theta _{i}\left( v\right) \partial\theta_{j}\left( v\right) }\right\vert ^{m}\right\} < \infty

    and

    \ E\left\{ \underset{\underline{\theta} \in\vartheta\left( \underline{\theta}_{0}\right) }{\sup}\left\vert \dfrac{1\ }{\sigma_{n,st+v}}\dfrac{\partial^{3}\sigma_{n,st+v}\left( \underline{\theta}\right) \ }{\partial\theta_{i}\left( v\right) \partial\theta_{j}\left( v\right) \partial\theta_{k}\left( v\right) }\right\vert ^{m}\right\} < \infty.

    We now prove \mathbf{(b)} , clearly, \nabla_{\underline{\theta} }l_{t}\left(\underline{\theta}_{0}\right) is SPS and measurable with respect to \mathcal{F}_{t} = \sigma\left(\underline{\eta}_{n}, n\leq t\right). The property stated in proof \mathbf{(a)} ensures the existence of matrices U and K in Theorem 4. Moreover, it can be seen that \forall\underline{\lambda}\in\mathbb{R}^{l} , the sequence \left(\underline{\lambda}^{\prime}\nabla_{\underline{\theta}}l_{t}\left(\underline{\theta}_{0}\right), \mathcal{F}_{t}\right) is an SPS , PE , and square-integrable martingale difference. The C.L.T of Bllingsley implies

    \left( Ns\right) ^{-\frac{1}{2}}\sum\limits_{t = 1}^{N}\sum\nolimits_{v = 1} ^{s}\nabla_{\underline{\theta}}l_{st+v}\left( \underline{\theta}_{0}\right) \ \leadsto\mathcal{N}\left( \underline{O},U\right) \ as\; N\rightarrow \infty.



    [1] G. Avalos, I. Lasiecka, Exponential stability of a thermoelastic system without mechanical dissipation, Rend. Istit. Mat. Univ. Trieste, 28 (1997), 1–28.
    [2] S. Aydinlik, A. Kiris, W. Sumelka, Nonlocal vibration analysis of microstretch plates in the framework of space-fractional mechanics-theory and validation, Eur. Phys. J. Plus, 136 (2021), 1–17. doi: 10.1140/epjp/s13360-021-01110-x. doi: 10.1140/epjp/s13360-021-01110-x
    [3] W. H. Chen, Cauchy problems for thermoelastic plate equations with different damping mechanisms, Commun. Math. Sci., 18 (2020), 429–457. doi: 10.4310/cms.2020.v18.n2.a7. doi: 10.4310/cms.2020.v18.n2.a7
    [4] W. H. Chen, Decay properties and asymptotic profiles for elastic waves with Kelvin-Voigt damping in 2D, Asymptotic Anal., 117 (2019), 113–140. doi: 10.3233/ASY-191548. doi: 10.3233/ASY-191548
    [5] W. H. Chen, Dissipative structure and diffusion phenomena for doubly dissipative elastic waves in two space dimensions, J. Math. Anal. Appl., 486 (2020), 123922. doi: 10.1016/j.jmaa.2020.123922. doi: 10.1016/j.jmaa.2020.123922
    [6] W. H. Chen, M. Reissig, Weakly coupled systems of semilinear elastic waves with different damping mechanisms in 3D, Math. Method. Appl. Sci., 42 (2019), 667–709. doi: 10.1002/mma.5370. doi: 10.1002/mma.5370
    [7] G. Chen, D. L. Russell, A mathematical model for linear elastic systems with structural damping, Q. Appl. Math., 39 (1981), 433–454. doi: 10.1090/QAM/644099. doi: 10.1090/QAM/644099
    [8] S. P. Chen, R. Triggiani, Proof of extensions of two conjectures on structural damping for elastic systems, Pac. J. Math., 136 (1989), 15–55. doi: 10.2140/pjm.1989.136.15. doi: 10.2140/pjm.1989.136.15
    [9] S. P. Chen, R. Triggiani, Gevrey class semigroups arising from elastic systems with gentle dissipation: The case 0 < \alpha < \frac{1}{2}, Proc. Amer. Math. Soc., 110 (1990), 401–415. doi: 10.1090/S0002-9939-1990-1021208-4. doi: 10.1090/S0002-9939-1990-1021208-4
    [10] F. Dell'Oro, J. E. Muñoz Rivera, V. Pata, Stability properties of an abstract system with applications to linear thermoelastic plates, J. Evol. Equ., 13 (2013), 777–794. doi: 10.1007/s00028-013-0202-6. doi: 10.1007/s00028-013-0202-6
    [11] R. Denk, R. Racke, L^p-resolvent estimates and time decay for generalized thermoelastic plate equations, Electron. J. Differ. Equ., 2006 (2006), 1–16. doi: 10.1142/9789812772992_0015. doi: 10.1142/9789812772992_0015
    [12] R. Denk, R. Racke, Y. Shibata, L_p theory for the linear thermoelastic plate equations in bounded and exterior domains, Adv. Differ. Equ., 14 (2009), 685–715.
    [13] R. Denk, R. Racke, Y. Shibata, Local energy decay estimate of solutions to the thermoelastic plate equations in two- and three-dimensional exterior domains, Z. Anal. Anwend., 29 (2010), 21–62. doi: 10.4171/ZAA/1396. doi: 10.4171/ZAA/1396
    [14] J. H. Hao, Z. Y. Liu, Stability of an abstract system of coupled hyperbolic and parabolic equations, Z. Angew. Math. Phys., 64 (2013), 1145–1159. doi: 10.1007/s00033-012-0274-0. doi: 10.1007/s00033-012-0274-0
    [15] J. H. Hao, Z. Y. Liu, J. M. Yong, Regularity analysis for an abstract system of coupled hyperbolic and parabolic equations, J. Differ. Equ., 259 (2015), 4763–4798. doi: 10.1016/j.jde.2015.06.010. doi: 10.1016/j.jde.2015.06.010
    [16] F. Huang, On the mathematical model for linear elastic systems with analytic damping, SIAM J. Control Optim., 26 (1988), 714–724. doi: 10.1137/0326041. doi: 10.1137/0326041
    [17] F. Huang, K. Liu, Holomorphic property and exponential stability of the semigroup associated with linear elastic systems with damping, Ann. Differ. Equ., 4 (1988), 411–424.
    [18] J. U. Kim, On the energy decay of a linear thermoelastic bar and plate, SIAM J. Math. Anal., 23 (1992), 889–899. doi: 10.1137/0523047. doi: 10.1137/0523047
    [19] I. Lasiecka, R. Triggiani, Two direct proofs on the analyticity of the s.c. semigroup arising in abstract thermo-elastic equations, Adv. Differ. Equ., 3 (1998), 387–416.
    [20] I. Lasiecka, R. Triggiani, Analyticity, and lack thereof, of thermo-elastic semigroups, ESAIM: Proc., 4 (1998), 199–222. doi: 10.1051/proc:1998029.
    [21] I. Lasiecka, R. Triggiani, Analyticity of thermo-elastic semigroups with coupled hinged/Neumann B.C., Abstr. Appl. Anal., 3 (1998), 153–169. doi: 10.1155/S1085337598000487. doi: 10.1155/S1085337598000487
    [22] I. Lasiecka, R. Triggiani, Analyticity of thermo-elastic semigroups with free boundary conditions, Ann. Scuola Norm.-Sci., 27 (1998), 457–482.
    [23] Y. Liu, W. Chen, Asymptotic profifiles of solutions for regularity-loss-type generalized thermoelastic plate equations and their applications, Z. Angew. Math. Phys., 71 (2020), 1–26. doi: 10.1007/s00033-020-1283-z. doi: 10.1007/s00033-020-1283-z
    [24] K. Liu, Z. Liu, Exponential stability and analyticity of abstract linear thermoelastic systems, Z. Angew. Math. Phys., 48 (1997), 885–904. doi: 10.1007/s000330050071. doi: 10.1007/s000330050071
    [25] Z. Liu, J. Yong, Qualitative properties of certain C_0 semigroups arising in elastic systems with various dampings, Adv. Differ. Equ., 3 (1998), 643–686. doi: 10.1016/S1474-6670(17)56476-9. doi: 10.1016/S1474-6670(17)56476-9
    [26] K. Jachmann, A unified treatment of models of thermoelasticity, TU Bergakademie Freiberg, 2008.
    [27] J. E. Muñoz Rivera, R. Racke, Smoothing properties, decay, and global existence of solutions to nonlinear coupled systems of thermoelastic type, SIAM J. Math. Anal., 26 (1995), 1547–1563. doi: 10.1137/S0036142993255058. doi: 10.1137/S0036142993255058
    [28] J. E. Muñoz Rivera, R. Racke, Large solutions and smoothing properties for nonlinear thermoelastic systems, J. Differ. Equ., 127 (1996), 454–483. doi: 10.1006/jdeq.1996.0078. doi: 10.1006/jdeq.1996.0078
    [29] S. Patnaik, S. Sidhardh, F. Semperlotti, Fractional-order models for the static and dynamic analysis of nonlocal plates, Commun. Nonlinear Sci., 95 (2021), 105601. doi: 10.1016/j.cnsns.2020.105601. doi: 10.1016/j.cnsns.2020.105601
    [30] S. Patnaik, S. Sidhardh, F. Semperlotti, Nonlinear thermoelastic fractional-order model of nonlocal plates: Application to postbuckling and bending response, Thin-Wall. Struct., 164 (2021), 107809. doi: 10.1016/j.tws.2021.107809. doi: 10.1016/j.tws.2021.107809
    [31] R. Racke, Y. Ueda, Dissipative structures for thermoelastic plate equations in \mathbb{R}^n, Adv. Differ. Equ., 21 (2016), 610–630.
    [32] R. Racke, Y. Ueda, Nonlinear thermoelastic plate equations-Global existence and decay rates for the Cauchy problem, J. Differ. Equ., 263 (2017), 8138–8177. doi: 10.1016/j.jde.2017.08.036. doi: 10.1016/j.jde.2017.08.036
    [33] M. Reissig, Structurally damped elastic waves in 2D, Math. Method. Appl. Sci., 39 (2016), 4618–4628. doi: 10.1002/mma.3888. doi: 10.1002/mma.3888
    [34] M. Reissig, Y. G. Wang, Cauchy problems for linear thermoelastic systems of type III in one space variable, Math. Method. Appl. Sci., 28 (2005), 1359–1381. doi: 10.1002/mma.619. doi: 10.1002/mma.619
    [35] B. Said-Houari, Decay properties of linear thermoelastic plates: Cattaneo versus Fourier law, Appl. Anal., 92 (2013), 424–440. doi: 10.1080/00036811.2011.625015. doi: 10.1080/00036811.2011.625015
    [36] Y. G. Wang, L. Yang, L^p-L^q decay estimates for Cauchy problems of linear thermoelastic systems with second sound in three dimensions, P. Roy. Soc. Edinb. A, 136 (2006), 189–207. doi: 10.1017/s0308210500004510. doi: 10.1017/s0308210500004510
    [37] J. Xu, N. Mori, S. Kawashima, L^p-L^q-L^r estimates and minimal decay regularity for compressible Euler-Maxwell equations, J. Math. Pures Appl., 104 (2015), 965–981. doi: 10.1016/j.matpur.2015.07.001. doi: 10.1016/j.matpur.2015.07.001
    [38] K. Yagdjian, The Cauchy problem for hyperbolic operators. Multiple characteristics. Micro-local approach (Mathematical Topics 12), Berlin: Akademie Verlag, 1997.
    [39] H. Yu, G. J. Xu, Z. Q. Zheng, Transmission characteristics of terahertz waves propagation in magnetized plasma using the WKB method, Optik, 188 (2019), 244–250. doi: 10.1016/j.ijleo.2019.05.061. doi: 10.1016/j.ijleo.2019.05.061
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2402) PDF downloads(111) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog