Loading [MathJax]/jax/output/SVG/jax.js
Research article

Least squares type estimations for discretely observed nonergodic Gaussian Ornstein-Uhlenbeck processes of the second kind

  • Received: 21 May 2021 Accepted: 10 October 2021 Published: 20 October 2021
  • MSC : 62F12, 60G22

  • We consider the nonergodic Gaussian Ornstein-Uhlenbeck processes of the second kind defined by dXt=θXtdt+dY(1)t,t0,X0=0 with an unknown parameter θ>0, where dY(1)t=etdGat and {Gt,t0} is a mean zero Gaussian process with the self-similar index γ(12,1) and at=γetγ. Based on the discrete observations {Xti:ti=iΔn,i=0,1,,n}, two least squares type estimators ˆθn and ˜θn of θ are constructed and proved to be strongly consistent and rate consistent. We apply our results to the cases such as fractional Brownian motion, sub-fractional Brownian motion, bifractional Brownian motion and sub-bifractional Brownian motion. Moreover, the numerical simulations confirm the theoretical results.

    Citation: Huantian Xie, Nenghui Kuang. Least squares type estimations for discretely observed nonergodic Gaussian Ornstein-Uhlenbeck processes of the second kind[J]. AIMS Mathematics, 2022, 7(1): 1095-1114. doi: 10.3934/math.2022065

    Related Papers:

    [1] Abdulaziz Alsenafi, Mishari Al-Foraih, Khalifa Es-Sebaiy . Least squares estimation for non-ergodic weighted fractional Ornstein-Uhlenbeck process of general parameters. AIMS Mathematics, 2021, 6(11): 12780-12794. doi: 10.3934/math.2021738
    [2] Chunhao Cai, Min Zhang . A note on inference for the mixed fractional Ornstein-Uhlenbeck process with drift. AIMS Mathematics, 2021, 6(6): 6439-6453. doi: 10.3934/math.2021378
    [3] Yajun Song, Ruyue Hu, Yifan Wu, Xiaohui Ai . Analysis of a stochastic two-species Schoener's competitive model with Lévy jumps and Ornstein–Uhlenbeck process. AIMS Mathematics, 2024, 9(5): 12239-12258. doi: 10.3934/math.2024598
    [4] Jiangrui Ding, Chao Wei . Parameter estimation for discretely observed Cox–Ingersoll–Ross model driven by fractional Lévy processes. AIMS Mathematics, 2023, 8(5): 12168-12184. doi: 10.3934/math.2023613
    [5] Jingwen Cui, Hao Liu, Xiaohui Ai . Analysis of a stochastic fear effect predator-prey system with Crowley-Martin functional response and the Ornstein-Uhlenbeck process. AIMS Mathematics, 2024, 9(12): 34981-35003. doi: 10.3934/math.20241665
    [6] Shaoling Zhou, Huixin Chai, Xiaosheng Wang . Barrier option pricing with floating interest rate based on uncertain exponential Ornstein–Uhlenbeck model. AIMS Mathematics, 2024, 9(9): 25809-25833. doi: 10.3934/math.20241261
    [7] Khaled M. Alqahtani, Mahmoud El-Morshedy, Hend S. Shahen, Mohamed S. Eliwa . A discrete extension of the Burr-Hatke distribution: Generalized hypergeometric functions, different inference techniques, simulation ranking with modeling and analysis of sustainable count data. AIMS Mathematics, 2024, 9(4): 9394-9418. doi: 10.3934/math.2024458
    [8] Chunting Ji, Hui Liu, Jie Xin . Random attractors of the stochastic extended Brusselator system with a multiplicative noise. AIMS Mathematics, 2020, 5(4): 3584-3611. doi: 10.3934/math.2020233
    [9] Ruyue Hu, Chi Han, Yifan Wu, Xiaohui Ai . Analysis of a stochastic Leslie-Gower three-species food chain system with Holling-II functional response and Ornstein-Uhlenbeck process. AIMS Mathematics, 2024, 9(7): 18910-18928. doi: 10.3934/math.2024920
    [10] Mingying Pan, Xiangchu Feng . Application of Fisher information to CMOS noise estimation. AIMS Mathematics, 2023, 8(6): 14522-14540. doi: 10.3934/math.2023742
  • We consider the nonergodic Gaussian Ornstein-Uhlenbeck processes of the second kind defined by dXt=θXtdt+dY(1)t,t0,X0=0 with an unknown parameter θ>0, where dY(1)t=etdGat and {Gt,t0} is a mean zero Gaussian process with the self-similar index γ(12,1) and at=γetγ. Based on the discrete observations {Xti:ti=iΔn,i=0,1,,n}, two least squares type estimators ˆθn and ˜θn of θ are constructed and proved to be strongly consistent and rate consistent. We apply our results to the cases such as fractional Brownian motion, sub-fractional Brownian motion, bifractional Brownian motion and sub-bifractional Brownian motion. Moreover, the numerical simulations confirm the theoretical results.



    Recently, the study of fractional Ornstein-Uhlenbeck processes of the second kind (FOUSK) has attracted interest. For example, Azmoodeh and Morlanes (2013) studied the drift parameter estimation of FOUSK based on continuous observations in the ergodic case. Azmoodeh and Viitasaari (2015) considered the drift parameter estimation of FOUSK based on discrete observations in the ergodic case. For the nonergodic case, the drift parameter estimations of FOUSK based on continuous and discrete observations were studied in EI Onsy et al. (2017) and EI Onsy et al. (2018), respectively. Balde et al. (2018) investigated the infinite-dimensional version of FOUSK. Yu et al. (2017) studied the problem of parameter estimation for Ornstein-Uhlenbeck processes of the second kind driven by α-stable Lévy motions, based on continuous and discrete observations, respectively. Es-Sebaiy et al. (2019) considered least squares type estimations for discretely observed nonergodic Gaussian Ornstein-Uhlenbeck processes.

    Motivated by all these studies, in this paper, we will consider the nonergodic Gaussian Ornstein-Uhlenbeck processes of the second kind (GOUSK) defined by

    dXt=θXtdt+dY(1)t,t0,X0=0, (1.1)

    with an unknown parameter θ>0, where dY(1)t=etdGat and G={Gt,t0} is a mean zero Gaussian process with the self-similar index γ(12,1) and at=γetγ. Based on the discrete observations {Xti:ti=iΔn,i=0,1,,n}, we construct two least squares type estimators ˆθn and ˜θn of θ:

    ˆθn=ni=1Xti1(XtiXti1)Δnni=1X2ti1 (1.2)

    and

    ˜θn=X2Tn2Δnni=1X2ti1, (1.3)

    where Tn=nΔn denotes the length of the 'observation window'.

    Let the covariance function of Gaussian process G be R(t,s)=E(GtGs),t0,s0. Denote

    ϕ(t,s)=2R(t,s)ts. (1.4)

    Assume that

    ϕ(t,s)c(γ)|ts|2γ2, (1.5)

    where c(γ) is a positive constant depending on γ, and 12<γ<1 is self-similar index of G.

    For the assumption (1.5), many self-similar Gaussian processes satisfy the condition, such as fractional Brownian motion, sub-fractional Brownian motion, bifractional Brownian motion and sub-bifractional Brownian motion. Therefore our main results hold for the Gaussian processes mentioned above.

    Let {Zn} be a sequence of random variables. We say {Zn} is tight (or bounded in probability) if for every ϵ>0, there exists Mϵ>0, such that

    P(|Zn|>Mϵ)<ϵ    for all  n.

    Now we state our main results as follows.

    Theorem 1.1. Assume that (1.5) holds and 12<γ<1, and θ>0,Δn0 and nΔ1+βn for some β>0 as n. Then, we have, as n,

    ˆθna.s.θ, (1.6)

    and

    ˜θna.s.θ. (1.7)

    Theorem 1.2. Assume that (1.5) holds and 12<γ<1, and θ>0,Δn0 and nΔ1+βn for some β>0 as n. Then, we have

    (1) for any q0,

    ΔqneθTn(ˆθnθ)  is not tight, (1.8)

    (2) if nΔ3n0 as n, then

    Tn(ˆθnθ)  is tight. (1.9)

    Theorem 1.3. Assume that (1.5) holds and 12<γ<1, and θ>0,Δn0 and nΔ1+βn for some β>0 as n. Then, we have

    (1) for any q0,

    ΔqneθTn(˜θnθ)  is not tight, (1.10)

    (2) if nΔ3n0 as n, then

    Tn(˜θnθ)  is tight. (1.11)

    Remark. We study Gaussian Ornstein-Uhlenbeck processes of the second kind in this paper. We know they are a subset of a much larger class Barndorff-Nielsen and Shephard type Ornstein-Uhlenbeck processes (see Barndorff-Nielsen (2001) and Barndorff-Nielsen and Shephard (2001)). In the future, we will extend our results to more general Ornstein-Uhlenbeck type models (see Salmon and SenGupta (2021), Issaka and SenGupta (2017) and Roberts and SenGupta (2020)).

    We have organized our paper as follows: In Sect.2 we present some preliminaries for the Gaussian process G and main lemmas. Sect.3 is devoted to the proofs of Theorems 1.1–1.3. In Sect.4 we apply our results to the cases such as fractional Brownian motion, sub-fractional Brownian motion, bifractional Brownian motion and sub-bifractional Brownian motion, while Sect.5 contains numerical simulations for four fractional Gaussian processes.

    In this section, we firstly recall some elements of the Malliavin calculus. We refer to Nualart (2006) for detailed account these notions(see Moshrefi-Torbati and Hammond (1998), Meerschaert et al. (2017), Ei-Nabulsi (2012, 2015, 2017), and Ei-Nabulsi and Golmankhaneh (2021)). Let H be a real separable Hilbert space associated with the Gaussian process G, which is defined by the closure of the linear space ε generated by the indicator functions {I[0,t],t[0,T]} with respect to the scalar product

    I[0,t],I[0,s]H=R(t,s).

    We know that the covariance of G can be written as

    R(t,s)=t0s0ϕ(u,v)dudv, (2.1)

    where ϕ(u,v) is defined by (1.4).

    We can find a linear space of functions contained in H in the following way. Let |H| be the linear space of measurable functions φ on [0,T] such that

    ||φ||2|H|=T0T0|φ(u)||φ(v)|ϕ(u,v)dudv<.

    It is not difficult to show that |H| is a Banach space with the norm |||||H| and ε is dense in |H|.

    Moreover, for all φ,ψ|H|, it can be proved that

    E(T0φ(u)dGuT0ψ(v)dGv)=T0T0φ(u)ψ(v)ϕ(u,v)dudv. (2.2)

    For every q1, let Hq be the qth Wiener chaos of G, namely, the closed linear subspace of L2(Ω) generated by the random variables {Hq(G(h)),hH,||h||H=1}, where Hq is the qth Hermite polynomial defined as Hq(x)=(1)qex22dqdxq(ex22). The mapping Iq(hq)=Hq(G(h)) provides a linear isometry between the symmetric tensor product Hq (equipped with the modified norm ||||Hq=q!||||Hq) and Hq. Specifically, for all f,gHq and q1, one has

    E[Iq(f)Iq(g)]=q!f,gHq. (2.3)

    For the multiple stochastic integral Iq(f), it has the following property: for any p2,

    (E[|Iq(f)|p])1/pc(p,q)(E[|Iq(f)|2])1/2, (2.4)

    where c(p,q) is a positive constant only depending on p and q.

    It is easy to obtain the solution of (1.1):

    Xt=eθtt0eθsdY(1)s,    t0, (2.5)

    where the integral with respect to Y(1)s is a Young integral (see Young (1936)).

    Denote

    ηt=t0eθsdY(1)s,    t0, (2.6)

    then,

    Xt=eθtηt. (2.7)

    By (2.6) and dY(1)s=esdGas, as=γesγ and let as=u, we get

    ηt=t0e(θ+1)sdGas               
    =γ(θ+1)γata0u(θ+1)γdGu. (2.8)

    By (2.2) and 12<γ<1, we have, for 0s<t,

    E[(ηtηs)2]=γ2(θ+1)γatasatas(uv)(θ+1)γϕ(u,v)dudv. (2.9)

    In order to prove Theorems 1.1-1.3, we need the following some lemmas.

    Lemma 2.1. Let η={ηt,t0} be given by (2.6). Assume that (1.5) holds and 12<γ<1. Then,

    (1) For all ϵ(0,γ), the process η has a modification with (γϵ)-Hölder continuous paths, still denoted η in the sequel.

    (2) As t,

    ηtη:=γ(θ+1)γa0u(θ+1)γdGu, (2.10)

    almost surely and in L2(Ω).

    Proof. Since the proof is similar to that of Lemma 2.2 in EI Onsy et al. (2017), we omit the details.

    Lemma 2.2. Let l>0 and {Zn}nN be a sequence of random variables. Suppose that, for every p1, there exists a constant cp>0 such that, for all nN,

    (E[|Zn|p])1/pcpnl.

    Then, for all ϵ>0, there exists a random variable ξϵ such that, for any nN,

    |Zn|ξϵnl+ϵ    a.s.,

    moreover, E[|ξϵ|p]< for all p1.

    Proof. See the proof in Kloeden and Neuenkirch (2007).

    Lemma 2.3. Let

    Rn=n1i=1e2θ(ni)Δn(η2tiη2ti1), (2.11)

    and

    Sn=Δnni=1X2ti1. (2.12)

    Then, we have

    e2θTnSn=Δne2θΔn1(η2tn1Rn). (2.13)

    Assume that (1.5) holds and 12<γ<1, and θ>0,Δn0 and nΔ1+βn for some β>0 as n, then,

    Rna.s.0. (2.14)

    In particular, as n, we have

    e2θTnSna.s.η22θ, (2.15)

    where η is defined by (2.10).

    Proof. By (2.7), we obtain

    e2θTnSn=e2θTnΔnni=1e2θ(i1)Δnη2ti1                                           
    =Δne2θΔn1ni=1e2θ(ni+1)Δn(e2θΔn1)η2ti1        
    =Δne2θΔn1ni=1(e2θ(ni)Δne2θ(ni+1)Δn)η2ti1 
    =Δne2θΔn1[η2tn1n1i=1e2θ(ni)Δn(η2tiη2ti1)].
    =Δne2θΔn1(η2tn1Rn),                                    

    which shows that (2.13) holds.

    Using (2.9) and (1.5) and making the change of variables r=uati1, s=vati1, we get

    E[(ηtiηti1)2]=γ2(θ+1)γatiati1atiati1(uv)(θ+1)γϕ(u,v)dudv                                       
    c(γ)γ2(θ+1)γatiati1atiati1(uv)(θ+1)γ|uv|2γ2dudv      
    =c(γ)γ2γe2θ(i1)ΔnΔnγ1Δnγ1(rs)(θ+1)γ|rs|2γ2drds
    c(γ)γ2γe2θ(i1)ΔnΔnγ1Δnγ1|rs|2γ2drds              
    =c(γ)γ2γe2θ(i1)Δnγ(2γ1)(eΔnγ1)2γ.                              (2.16)

    By (2.10) and η is Gaussian, we also obtain, for every p1,

    (E[|η2tiη2ti1|p])1/pc(p)(E[(ηtiηti1)2])1/2, (2.17)

    where c(p) is a constant depending on p. In fact, we have, by Cauchy-Schwarz inequality,

    (E[|η2tiη2ti1|p])1/p=(E[|ηti+ηti1|p|ηtiηti1|p])1/p                      
                               {[E|ηti+ηti1|2p]1/2[E|ηtiηti1|2p]1/2}1/p
                    [E|2η|2p]1/(2p)[E|ηtiηti1|2p]1/(2p)
    c(p)(E[(ηtiηti1)2])1/2,

    the last inequality comes from the fact: if ξ is Gaussian and p2, then

    [E|ξ|p]1/pc(p)[E|ξ|2]1/2.

    By (2.17) and (2.16) and the Minkowski inequality, we deduce that

    (E[|Rn|p])1/pn1i=1e2θ(ni)Δn(E[|η2tiη2ti1|p])1/p                                    
    c(p)c(γ)γγγ(2γ1)eθnΔn(eΔnγ1)γn1i=1eθ(ni1)Δn  
    =c(p)c(γ)γγγ(2γ1)eθnΔn(eΔnγ1)γ1eθ(n1)Δn1eθΔn    
    c(p)c(γ)γγγ(2γ1)eθnΔn(eΔnγ1)γ11eθΔn          
    =c(p)c(γ)θγ(2γ1)(Δn)γ1eθnΔn(eΔnγ1Δnγ)γeθΔneθΔn1θΔn
    c(p)c(γ)cθγ(2γ1)(Δn)γ1eθnΔn,                            (2.18)

    where c is a constant, and the last inequality comes from the fact: as x0,

    ex1x1.

    Note that for any α>0, as n,

    (Δn)γ1eθnΔn=o(nα). (2.19)

    In fact,

    nα(Δn)γ1eθTn=Tα+α+1γβneθTn(nΔ1+βn)α+1γβ. (2.20)

    Since nΔ1+βn for some β>0 as n, we have

    (nΔ1+βn)α+1γβ. (2.21)

    In addition,

    limnTα+α+1γβneθTn=0. (2.22)

    Thus (2.19) is obtained from (2.20)–(2.22).

    By (2.19), we have, for any δ>0,

    (Δn)γ1eθTn=nαδ.

    Hence (2.18) becomes

    (E[|Rn|p])1/pc(p)c(γ)cθγ(2γ1)nαδ.

    By Lemma 2.2, there exists a random variable ξα such that

    |Rn|ξαnα,    a.s.

    for all nN. Moreover, E|ξα|p< for all p1. Therefore (2.14) holds.

    Finally, we can get (2.15) using (2.13), (2.14), (2.10) and limnΔne2θΔn1=12θ. The proof of Lemma 2.3 is finished.

    In this section, we will give proofs of Theorems 1.1-1.3.

    By (2.7), we can rewrite ˆθn and ˜θn as follows, respectively,

    ˆθn=ni=1eθtiηtiXti1ni=1X2ti1Sn                                                
    =ni=1eθti(ηtiηti1)Xti1+(ni=1eθtiηti1Xti1ni=1X2ti1)Sn
    =ni=1eθti(ηtiηti1)Xti1+(eθΔn1)ni=1X2ti1)Sn                
    =VnSn+eθΔn1Δn,                                                                (3.1)

    where

    Vn=ni=1eθti(ηtiηti1)Xti1,

    and

    ˜θn=η2Tn2e2θTnSn. (3.2)

    Proof of Theorem 1.1. For (1.6), according to (2.15) and (3.1), it suffices to prove that, as n,

    e2θTnVna.s.0. (3.3)

    By the Minkowski inequality and (2.16), we have

    [E|e2θTnVn|2]1/2=e2θTn{E[ni=1eθti(ηtiηti1)Xti1]2}1/2                                                
    e2θTnni=1eθiΔn(EX2ti1)1/2[E(ηtiηti1)2]1/2                
    e2θTnc(γ)γγγ(2γ1)eθΔn(eΔnγ1)γni=1(EX2ti1)1/2            
    e2θTnc(γ)γγγ(2γ1)eθΔn(eΔnγ1)γ(Eη2)1/2ni=1eθ(i1)Δn   
    =c(γ)γγγ(2γ1)eθΔn(Eη2)1/2(eΔnγ1)γe2θTn1eθTn1eθΔn        
    =c(γ)γγγ(2γ1)eθΔn(Eη2)1/2(eΔnγ1)γeθTn1eθTneθΔn1        
    c(γ)γγγ(2γ1)eθΔn(Eη2)1/2(eΔnγ1)γeθTn1eθΔn1        
    =c(γ)θγ(2γ1)(Eη2)1/2(Δn)γ1(eΔnγ1Δnγ)γeθTneθΔneθΔn1θΔn
    c(γ)cθγ(2γ1)(Eη2)1/2(Δn)γ1eθTn,                             (3.4)

    where c is a generic constant, and the last inequality also comes from the fact: as x0,

    ex1x1.

    Using similar arguments as the proof of (2.14), we can obtain (3.3). Thus (1.6) holds.

    We can easily obtain (1.7) from (3.2), (2.15) and limtηt=η a.s., and this completes the proof of Theorem 1.1.

    Proof of Theorem 1.2. (1) Firstly, we consider the case q12. By (3.1), we get

    ΔqneθTn(ˆθnθ)=ΔqneθTn(VnSn+eθΔn1Δnθ)                                       
    =ΔqneθTnVne2θTnSn+Δq1neθTn(eθΔn1θΔn). (3.5)

    Since eθΔn1θΔnθ22Δ2n as Δn0, we obtain

    limnΔq1neθTn(eθΔn1θΔn)=limnθ22Δq+1neθTn                                                         
    =limnθ22(nΔ1+βn)q+1βeθTnTq+1βn.

    Because nΔ1+βn for some β>0 as n and

    limneθTnTq+1βn=.

    Then, we have

    limnΔq1neθTn(eθΔn1θΔn)=. (3.6)

    On the other hand, by (3.4), we get, as n,

    E|ΔqneθTnVn|ΔqneθTn(EV2n)1/2
                                        c(γ)cθγ(2γ1)(Eη2)1/2Δq+γ1n
    0, (3.7)

    since γ>12. Consequently, by (3.5)–(3.7) and (2.15), for any q12, ΔqneθTn(ˆθnθ) is not tight.

    For the case 0q<12, note that

    ΔqneθTn(ˆθnθ)=Δq12n(Δ12neθTn(ˆθnθ)),

    and limnΔq12n= and the previous case q=12. The proof of (1.8) is finished.

    (2) By (3.1), it is obvious that

    Tn(ˆθnθ)=nΔ3neθΔn1θΔnΔ2n+Tne2θTnVne2θTnSn. (3.8)

    Since nΔ3n0 and eθΔn1θΔnΔ2nθ22, we have, as n,

    nΔ3neθΔn1θΔnΔ2n0. (3.9)

    Furthermore, by (3.4), we get, as n,

    E|Tne2θTnVn|Tn[E|e2θTnVn|2]1/2                              
                  c(γ)cθγ(2γ1)(Eη2)1/2Δγ1nTneθTn
               =c(γ)cθγ(2γ1)(Eη2)1/2T12+1γβneθTn(nΔ1+βn)1γβ
    0.                                  (3.10)

    By (3.8)–(3.10) and (2.15), we deduce (1.9). Hence, we complete the proof of Theorem 1.2.

    Proof of Theorem 1.3. (1) Fix q12, by (3.2) and (2.13), we obtain

    ΔqneθTn(˜θnθ)                                                                                               
    =ΔqneθTn(η2Tn2e2θTnSnθ)                                                                                  
    =ΔqneθTn2e2θTnSn[(η2tnη2tn1)+(12θΔne2θΔn1)η2tn12θ(e2θTnSnΔne2θΔn1η2tn1)]
    =ΔqneθTn2e2θTnSn[(η2tnη2tn1)+(12θΔne2θΔn1)η2tn1+(2θΔne2θΔn1)Rn].             (3.11)

    By (2.16), we have, as n,

    [E(ΔqneθTn(η2tnη2tn1))2]1/2Δqn[E(ηtn+ηtn1)4]1/4[E(eθTn(ηtnηtn1))4]1/4
                                    23[Eη2]1/2Δqn[E(eθTn(ηtnηtn1))2]1/2
                                     23c(γ)γ(2γ1)(Eη2)1/2eθΔnΔq+γn(eΔnγ1Δnγ)γ
    0.                        (3.12)

    We also get

    ΔqneθTn(12θΔne2θΔn1)=Δq+1neθTne2θΔn12θΔnΔ2nΔne2θΔn1           
                                       =(nΔ1+βn)q+1βeθTnTq+1βne2θΔn12θΔnΔ2nΔne2θΔn1
    ,  as  n,    (3.13)

    since limnnΔ1+βn= for some β>0, limneθTnTq+1βn=, limne2θΔn12θΔnΔ2n=2θ2 and limnΔne2θΔn1=12θ.

    Moreover, by (2.18), we deduce, as n,

    [E|ΔqneθTnRn|2]1/2=ΔqneθTn[E|Rn|2]1/2          
                      c(2)c(γ)cθγ(2γ1)Δq+γ1n
    0.       (3.14)

    where c(2) denotes c(p)=c(2) in (2.18). Combining (3.11)–(3.14) and (2.15), we conclude that for every q12, ΔqneθTn(˜θnθ) is not tight.

    For the case 0q<12, we obtain it similarly to the proof of (1) in Theorem 1.2. Hence (1.10) holds.

    (2) By (3.2) and (2.13), note that Tn=tn=nΔn, we can write

    Tn(˜θnθ)                                                                                    
    =Tn(η2Tn2e2θTnSnθ)                                                                  
    =Tn2e2θTnSn[(η2tnη2tn1)+(12θΔne2θΔn1)η2tn1+(2θΔne2θΔn1)Rn]. (3.15)

    Similarly to (3.12)–(3.14), we obtain, as n,

    [E(Tn(η2tnη2tn1))2]1/223c(γ)γ(2γ1)(Eη2)1/2TneθTneθΔnΔγn0, (3.16)
    Tn(12θΔne2θΔn1)=nΔ3ne2θΔn12θΔnΔ2nΔne2θΔn10, (3.17)

    and

    [E|TnRn|2]1/2c(2)c(γ)cθγ(2γ1)TneθTnΔγ1n                    
    =c(2)c(γ)cθγ(2γ1)T12+1γβneθTn(nΔ1+βn)1γβ
    0.                                  (3.18)

    By (3.15)–(3.18), we can easily get (1.11).

    Thus, we finish the proof of Theorem 1.3.

    This section is devoted to some examples of the Gaussian process G. For example fractional Brownian motion, sub-fractional Brownian motion, bifractional Brownian motion and sub-bifractional Brownian motion.

    The fractional Brownian motion (fBm) BH={BHt,t0} with Hurst parameter H(0,1) is defined as a centered Gaussian process starting from zero with covariance

    R(t,s)=E(BHtBHs)=12(t2H+s2H|ts|2H). (4.1)

    Note that, when H=12, B12 is a standard Brownian motion. By (4.1), we get

    ϕ(t,s)=2R(t,s)ts=H(2H1)|ts|2H2. (4.2)

    It is well-known that the self-similar index of fBm is H. Thus when 12<H<1, fBm satisfies the assumption (1.5). Hence Theorems 1.1–1.3 hold for the fBm BH(12<H<1), namely, corresponding to Theorems 4, 6, and 7 in EI Onsy et al. (2019).

    The sub-fractional Brownian motion (sfBm) SH={SHt,t0} with Hurst parameter H(0,1) is defined as a centered Gaussian process starting from zero with covariance

    R(t,s)=E(SHtSHs)=t2H+s2H12((t+s)2H+|ts|2H). (4.3)

    Note that, when H=12, S12 is a standard Brownian motion. The self-similar index for sfBm is also H. For more on sub-fractional Brownian motion, we can see Kuang and Xie (2015, 2017), Kuang and Liu (2015, 2018) and so on.

    By (4.3), we have

    ϕ(t,s)=2R(t,s)ts=H(2H1)[|ts|2H2(t+s)2H2]. (4.4)

    When 12<H<1, by (4.4), we get

    ϕ(t,s)H(2H1)|ts|2H2, (4.5)

    which shows that the assumption (1.5) holds for sfBm. Hence Theorems 1.1–1.3 hold for the sfBm SH(12<H<1).

    The bifractional Brownian motion (bfBm) BH,K={BH,Kt,t0} with parameters H(0,1) and K(0,1] is defined as a centered Gaussian process starting from zero with covariance

    R(t,s)=E(BH,KtBH,Ks)=12K((t2H+s2H)K|ts|2HK). (4.6)

    Note that, the case K=1 corresponds to the fBm with Hurst parameter H. The self-similar index of bfBm is HK. By (4.6), we obtain

    ϕ(t,s)=2R(t,s)ts=22KH2K(K1)(t2H+s2H)K2(ts)2H1                
    +21KHK(2HK1)|ts|2HK2. (4.7)

    Since K1, when 1<2HK<2, we have

    ϕ(t,s)21KHK(2HK1)|ts|2HK2. (4.8)

    The assumption (1.5) holds for the bfBm BH,K(1<2HK<2). Thus we also obtain Theorems 1.1-1.3 for the bfBm BH,K(1<2HK<2).

    El-Nouty and Journˊe (2013) introduced the process SH,K={SH,Kt,t0} with indices H(0,1) and K(0,1], named the sub-bifractional Brownian motion (sbfBm) and defined as follows:

    SH,Kt=12(2K)/2(BH,Kt+BH,Kt),

    where {BH,Kt,tR} is a bifractional Brownian motion (bfBm) with indices H(0,1) and K(0,1]. Clearly, the sbfBm is a centered Gaussian process such thatSH,K0=0, with probability 1, and Var(SH,Kt)=(2K22HK1)t2HK. Note that since (2H1)K1<K10, it follows that 2HK1<K. We can easily verify that SH,K is self-similar with index HK. When K=1, SH,1 is the sub-fractional Brownian motion (sfBm). Straightforward computations show that for all s,t0,

    R(t,s)=E(SH,KtSH,Ks)=(t2H+s2H)K12[(t+s)2HK+|ts|2HK], (4.9)

    and

    C1|ts|2HKE[(SH,KtSH,Ks)2]C2|ts|2HK, (4.10)

    where

    C1=min{2K1,2K22HK1},    C2=max{1,222HK1}.

    So SH,K has (HKϵ)-Hölder continuous paths for any ϵ(0,HK) thanks to Kolmogorov's continuity criterion. Kuang (2019) studied the collision local time of sub-bifractional Brownian Motions. Kuang and Li (2020) obtained Berry-Esséen bounds and proved the almost sure central limit theorem for the quadratic variation of the sub-bifractional Brownian motion.

    By (4.9), we have

    ϕ(t,s)=2R(t,s)ts=4H2K(K1)(t2H+s2H)K2(ts)2H1                
    +HK(2HK1)[|ts|2HK2(t+s)2HK2]. (4.11)

    Since K1, when 1<2HK<2, we have

    ϕ(t,s)HK(2HK1)|ts|2HK2, (4.12)

    which means the assumption (1.5) also holds for the sbfBm SH,K(1<2HK<2). Hence Theorems 1.1–1.3 also hold for the sbfBm SH,K(1<2HK<2).

    In this section, we firstly simulate the sample paths of the process X given by (1.1). By (2.7) and (2.8), we have

    Xt=eθtγ(θ+1)γata0u(θ+1)γdGu.

    Let ti=iΔn,i=0,1,,n,Xt0=0, then,

    Xti=eθtiγ(θ+1)γatia0u(θ+1)γdGu                                                                 
    =eθtiγ(θ+1)γ(ati1a0u(θ+1)γdGu+atiati1u(θ+1)γdGu)                       
    =eθΔnXti1+eθtiγ(θ+1)γatiati1u(θ+1)γdGu                                         
    eθΔnXti1+γ(θ+1)γeθiΔn(ati1+ati2)(θ+1)γ(GatiGati1)            
    =eθΔnXti1+γ(θ+1)γeθiΔn[γ2(eiΔnγ+e(i1)Δnγ)](θ+1)γ(GatiGati1), (5.1)

    for i=1,2,,n.

    Let

    Gati=Var(Gati)ξi,    i=0,1,,n, (5.2)

    where ξi(i=0,1,,n) are standard normal random variables.

    We know that

    ˆθn=ni=1Xti1(XtiXti1)Δnni=1X2ti1

    and

    ˜θn=X2Tn2Δnni=1X2ti1.

    Thus, we can obtain the simulations of ˆθn and ˜θn.

    Case 1: if G is fBm BH(12<H<1), then (5.1) and (5.2) become (5.3) and (5.4), respectively,

    Xti=eθΔnXti1+H(θ+1)HeθiΔn[H2(eiΔnH+e(i1)ΔnH)](θ+1)H(BHatiBHati1), (5.3)

    and

    BHati=(HeiΔnH)Hξi,    i=0,1,,n. (5.4)

    Case 2: if G is sfBm SH(12<H<1), then (5.1) and (5.2) become (5.5) and (5.6), respectively,

    Xti=eθΔnXti1+H(θ+1)HeθiΔn[H2(eiΔnH+e(i1)ΔnH)](θ+1)H(SHatiSHati1), (5.5)

    and

    SHati=222H1(HeiΔnH)Hξi,    i=0,1,,n. (5.6)

    Case 3: if G is bfBm BH,K(1<2HK<2), then (5.1) and (5.2) are replaced by (5.7) and (5.8), respectively,

    Xti=eθΔnXti1+(HK)(θ+1)HKeθiΔn[HK2(eiΔnHK+e(i1)ΔnHK)](θ+1)HK(BH,KatiBH,Kati1), (5.7)

    and

    BH,Kati=(HKeiΔnHK)HKξi,    i=0,1,,n. (5.8)

    Case 4: if G is sbfBm SH,K(1<2HK<2), then (5.1) and (5.2) become (5.9) and (5.10), respectively,

    Xti=eθΔnXti1+(HK)(θ+1)HKeθiΔn[HK2(eiΔnHK+e(i1)ΔnHK)](θ+1)HK(SH,KatiSH,Kati1), (5.9)

    and

    SH,Kati=2K22HK1(HKeiΔnHK)HKξi,    i=0,1,,n. (5.10)

    Now we take Δn=0.0002,n=2×105, and simulate 500 sample paths of X for different values of H,θ or H,K,θ. The results of simulations are summarized in Tables 14 respectively. From these results, it can be seen that the mean of both constructed parameter estimators are close to the true parameter values and the standard deviations of ˆθn and ˜θn are very small. Hence the numerical simulations confirm the theoretical research.

    Table 1.  Mean and standard deviation of two estimators for fBm.
    H=0.55 H=0.60 H=0.65 H=0.70
    θ=0.8
    Mean(ˆθn) 0.8000 0.8000 0.8000 0.8000
    SD(ˆθn) 8.9194e-08 2.6459e-07 1.4519e-07 5.6355e-07
    Mean(˜θn) 0.8001 0.8001 0.8001 0.8001
    SD(˜θn) 8.9197e-08 2.6461e-07 1.4520e-07 5.6338e-07
    θ=1.7
    Mean(ˆθn) 1.7001 1.7001 1.7001 1.7001
    SD(ˆθn) 1.0967e-14 3.1736e-14 1.2694e-14 1.2644e-14
    Mean(˜θn) 1.7003 1.7003 1.7003 1.7003
    SD(˜θn) 1.0865e-14 3.1832e-14 1.2580e-14 1.3038e-14
    θ=3.7
    Mean(ˆθn) 3.7007 3.7007 3.7007 3.7007
    SD(ˆθn) 1.1281e-14 1.0869e-14 1.1878e-14 1.1458e-14
    Mean(˜θn) 3.7014 3.7014 3.7014 3.7014
    SD(˜θn) 1.0194e-14 9.9964e-15 1.1395e-14 1.1431e-14

     | Show Table
    DownLoad: CSV
    Table 2.  Mean and standard deviation of two estimators for sfBm.
    H=0.55 H=0.60 H=0.65 H=0.70
    θ=0.8
    Mean(ˆθn) 0.7967 0.8000 0.8000 0.8000
    SD(ˆθn) 2.6371e-02 3.9579e-07 2.0294e-07 1.2505e-06
    Mean(˜θn) 0.8000 0.8001 0.8001 0.8001
    SD(˜θn) 2.0488e-04 3.9564e-07 2.0297e-07 1.2463e-06
    θ=1.7
    Mean(ˆθn) 1.7001 1.7001 1.7001 1.7001
    SD(ˆθn) 1.2324e-14 1.6913e-14 1.3377e-14 1.6702e-14
    Mean(˜θn) 1.7003 1.7003 1.7003 1.7003
    SD(˜θn) 1.1958e-14 1.6863e-14 1.3260e-14 1.6726e-14
    θ=3.7
    Mean(ˆθn) 3.7007 3.7007 3.7007 3.7007
    SD(ˆθn) 1.1491e-14 1.0355e-14 1.0807e-14 1.0539e-14
    Mean(˜θn) 3.7014 3.7014 3.7014 3.7014
    SD(˜θn) 1.1411e-14 1.0602e-14 1.0469e-14 1.0630e-14

     | Show Table
    DownLoad: CSV
    Table 3.  Mean and standard deviation of two estimators for bfBm.
    H=0.6  K=0.9 H=0.8  K=0.9 H=0.9  K=0.9
    θ=0.8
    Mean(ˆθn) 0.8000 0.8000 0.8000
    SD(ˆθn) 3.4198e-07 2.7941e-06 3.5679e-07
    Mean(˜θn) 0.8001 0.8001 0.8001
    SD(˜θn) 3.4207e-07 2.6092e-06 3.5678e-07
    θ=1.7
    Mean(ˆθn) 1.7001 1.7001 1.7001
    SD(ˆθn) 1.0361e-14 4.0730e-12 1.0888e-14
    Mean(˜θn) 1.7003 1.7003 1.7003
    SD(˜θn) 9.4681e-15 4.0742e-12 1.0035e-14
    θ=3.7
    Mean(ˆθn) 3.7007 3.7007 3.7007
    SD(ˆθn) 1.0713e-14 1.2432e-14 1.1264e-14
    Mean(˜θn) 3.7014 3.7014 3.7014
    SD(˜θn) 1.0130e-14 1.1353e-14 1.0026e-14

     | Show Table
    DownLoad: CSV
    Table 4.  Mean and standard deviation of two estimators for sbfBm.
    H=0.6  K=0.9 H=0.8  K=0.9 H=0.9  K=0.9
    θ=0.8
    Mean(ˆθn) 0.8000 0.8000 0.8000
    SD(ˆθn) 5.1295e-07 2.9938e-07 7.7103e-07
    Mean(˜θn) 0.8001 0.8001 0.8001
    SD(˜θn) 5.1320e-07 2.9930e-07 7.7119e-07
    θ=1.7
    Mean(ˆθn) 1.7001 1.7001 1.7001
    SD(ˆθn) 1.2144e-14 1.1284e-14 1.4808e-14
    Mean(˜θn) 1.7003 1.7003 1.7003
    SD(˜θn) 1.1640e-14 1.1001e-14 1.4746e-14
    θ=3.7
    Mean(ˆθn) 3.7007 3.7007 3.7007
    SD(ˆθn) 1.1830e-14 1.0794e-14 1.1455e-14
    Mean(˜θn) 3.7014 3.7014 3.7014
    SD(˜θn) 1.0983e-14 9.7612e-15 1.1214e-14

     | Show Table
    DownLoad: CSV

    EI Onsy et al. (2018) considered the parameter estimation for discretely observed nonergodic fractional Ornstein-Uhlenbeck process of the second kind. We extend their results to the case of Gaussian Ornstein-Uhlenbeck process of the second kind. We prove that two least squares type estimators are strongly consistent and rate consistent. Moreover, we give the numerical simulations which confirm the theoretical results. In the future, we will extend our results to more general Ornstein-Uhlenbeck type models such as Barndorff-Nielsen and Shephard type Ornstein-Uhlenbeck processes. On the other hand, we will consider the case of 0<γ<12 for the self-similar index γ of Gaussian process G.

    Nenghui Kuang was supported by the Natural Science Foundation of Hunan Province under Grant 2021JJ30233. Huantian Xie was supported in part by the NSF of Shandong Province (No.ZR2018LA008, 2019KJI003). The authors wish to thank both anonymous referees for careful reading of the previous versions of this paper and also their comments which improved the paper.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.



    [1] E. Azmoodeh, I. Morlanes, Drift parameter estimation for fractional Ornstein-Uhlenbeck process of the second kind, Statistics: A J. Theor. Appl. Stat., 49 (2015), 1–8. doi: 10.1080/02331888.2013.863888. doi: 10.1080/02331888.2013.863888
    [2] E. Azmoodeh, L. Viitasaari, Parameter estimation based on discrete observations of fractional Ornstein-Uhlenbeck process of the second kind, Stat. Infer. Stoch. Pro., 18 (2015), 205–227.
    [3] M. Balde, K. Es-Sebaiy, C. Tudor, Ergodicity and drift parameter estimation for infinite-dimensional fractional Ornstein-Uhlenbeck process of the second kind, Appl. Mat. Opt., 81 (2020), 785–814. doi: 10.1007/s00245-018-9519-4. doi: 10.1007/s00245-018-9519-4
    [4] O. Barndorff-Nielsen, Superposition of Ornstein-Uhlenbeck type processes, Theory Probab. Appl., 45 (2001), 175–194. doi: 10.1137/S0040585X97978166. doi: 10.1137/S0040585X97978166
    [5] O. Brndorff-Nielsen, N. Shephard, Non-Gaussian Ornstein-Uhlenbeck-based models and some of their uses in financial economics, J. Roy. Stat. Soc., 63 (2001), 167–241. doi: 10.1111/1467-9868.00282. doi: 10.1111/1467-9868.00282
    [6] R. A. El-Nabulsi, Glaeske-kilbas-saigo fractional integration and fractional dixmier trace, Acta Math. Vietnam., 37 (2012), 149–160.
    [7] R. A. El-Nabulsi, Fractional functional with two occurrences of integrals and asymptotic optimal change of drift in the black-scholes model, Acta Math. Vietnam., 40 (2015), 689–703. doi: 10.1007/s40306-014-0079-7. doi: 10.1007/s40306-014-0079-7
    [8] R. A. El-Nabulsi, Nonlocal-in-time kinetic energy in nonconservative fractional systems, disordered dynamics, jerk and snap and oscillatory motions in the rotating fluid tube, Int. J. Nonlinear Mech., 93 (2017), 65–81. doi: 10.1016/j.ijnonlinmec.2017.04.010. doi: 10.1016/j.ijnonlinmec.2017.04.010
    [9] R. A. El-Nabulsi, A. K. Golmankhaneh, Generalized heat diffusion equations with variable coefficients and their fractalization from the black-scholes equation, Comm. Theor. Phys., 73 (2021), 10–17.
    [10] B. EI Onsy, K. Es-Sebaiy, C. Tudor, Statistical analysis of the non-ergodic fractional Ornstein-Uhlenbeck process of the second kind, Commun. Stochastic Anal., 11 (2017), 65–81.
    [11] B. EI Onsy, K. Es-Sebaiy, D. Ndiaye, Parameter estimation for discretely observed non-ergodic fractional Ornstein-Uhlenbeck process of the second kind, Braz. J. Probab. Stat., 32 (2018), 545–558. doi: 10.1214/17-BJPS353. doi: 10.1214/17-BJPS353
    [12] K. Es-Sebaiy, F. Alazemi, M. Al-Foraih, Least squares type estimation for discretely observed non-ergodic Gaussian Ornstein-Uhlenbeck processes, Acta Math. Sci., 39 (2019), 989–1002.
    [13] A. Issaka, I. SenGupta, Analysis of variance based instruments for Ornstein-Uhlenbeck type models: swap and price index, Annals Financ., 13 (2017), 401–434. doi: 10.1007/s10436-017-0302-3. doi: 10.1007/s10436-017-0302-3
    [14] P. Kloeden, A. Neuenkirch, The pathwise convergence of approximation schemes for stochastic differential equations, LMS J. Comput. Math., 10 (2007), 235–253. doi: 10.1112/S1461157000001388. doi: 10.1112/S1461157000001388
    [15] N. Kuang, On the collision local time of sub-bifractional Brownian Motions, Adv. Math. (China), 48 (2019), 627–640.
    [16] N. Kuang, Y. Li, Berry-Esséen bounds and almost sure CLT for the quadratic variation of the sub-bifractional Brownian motion, Commun. Stat.-Simul. Comput., doi: 10.1080/03610918.2020.1740265. doi: 10.1080/03610918.2020.1740265
    [17] N. Kuang, B. Liu, Parameter estimations for the sub-fractional Brownian motion with drift at discrete observation, Braz. J. Probab. Stat., 29 (2015), 778–789. doi: 10.1214/14-BJPS246. doi: 10.1214/14-BJPS246
    [18] N. Kuang, B. Liu, Least squares estimator for α-sub-fractional bridges, Stat. Papers, 59 (2018), 893–912. doi: 10.1007/s00362-016-0795-2. doi: 10.1007/s00362-016-0795-2
    [19] N. Kuang, H. Xie, Maximum likelihood estimator for the sub-fractional Brownian motion approximated by a random walk, Ann. I. Stat. Math., 67 (2015), 75–91. doi: 10.1007/s10463-013-0439-4. doi: 10.1007/s10463-013-0439-4
    [20] N. Kuang, H. Xie, Asymptotic behavior of weighted cubic variation of sub-fractional brownian motion, Ann. I. Stat. Math., 46 (2017), 215–229. doi: 10.1080/03610918.2014.957849. doi: 10.1080/03610918.2014.957849
    [21] M. Moshrefi-Torbati, J. K. Hammond, Physical and geometrical interpretation of fractional operators, J. Franklin I., 335 (1998), 1077–1086. doi: 10.1016/S0016-0032(97)00048-3. doi: 10.1016/S0016-0032(97)00048-3
    [22] M. Meerschaert, E. Nane, P. Vellaisamy, The fractional poisson process and the inverse stable subordinator, Electron. J. Probab., 16 (2011), 1600–1620. doi: 10.1214/EJP.v16-920. doi: 10.1214/EJP.v16-920
    [23] D. Nualart, The Malliavin calculus and related topics, 2 Eds., Berlin: Springer-Verlag, 2006. doi: 10.1007/3-540-28329-3.
    [24] M. Roberts, I. SenGupta, Sequential hypothesis testing in machine learning, and crude oil price jump size detection, Appl. Math. Financ., 27 (2020), 374–395. doi: 10.1080/1350486X.2020.1859943. doi: 10.1080/1350486X.2020.1859943
    [25] N. Salmon, I. SenGupta, Fractional Barndorff-Nielsen and Shephard model: Applications in variance and volatility swaps, and hedging, arXiv: 2105.02325.
    [26] L. C. Young, An inequality of the hölder type, connected with stieltjes integration, Acta Math., 67 (1936), 251–282. doi: 10.1007/BF02401743. doi: 10.1007/BF02401743
    [27] Q. Yu, G. Shen, M. Cao, Parameter estimation for Ornstein-Uhlenbeck processes of the second kind driven by α-stable Lévy motions, Commun. Stat.-Theory M., 46 (2017), 10864–10878. doi: 10.1080/03610926.2016.1248786. doi: 10.1080/03610926.2016.1248786
  • This article has been cited by:

    1. Nenghui Kuang, Limits of sub-bifractional Brownian noises, 2023, 31, 2688-1594, 1240, 10.3934/era.2023063
    2. Nenghui Kuang, Huantian Xie, Derivative of self-intersection local time for the sub-bifractional Brownian motion, 2022, 7, 2473-6988, 10286, 10.3934/math.2022573
    3. Nenghui Kuang, Bingquan Liu, Renormalized self-intersection local time for sub-bifractional Brownian motion, 2022, 36, 0354-5180, 4023, 10.2298/FIL2212023K
    4. Jixia Wang, Lu Sun, Yu Miao, The estimations of drift parameters for the Gaussian Vasicek process with time-varying volatility, 2024, 53, 0361-0926, 8709, 10.1080/03610926.2023.2293650
    5. Nenghui Kuang, Huantian Xie, Least squares type estimators for the drift parameters in the sub-bifractional Vasicek processes, 2023, 26, 0219-0257, 10.1142/S0219025723500042
    6. Lili Bai, Chaopeng Guo, Jie Song, Cattle weight estimation model through readily photos, 2025, 143, 09521976, 109976, 10.1016/j.engappai.2024.109976
  • 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(2334) PDF downloads(81) Cited by(6)

Figures and Tables

Tables(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog