1.
Introduction and main 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
with an unknown parameter θ>0, where dY(1)t=e−tdGat and G={Gt,t≥0} 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:
and
where Tn=nΔn denotes the length of the 'observation window'.
Let the covariance function of Gaussian process G be R(t,s)=E(GtGs),t≥0,s≥0. Denote
Assume that
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
Now we state our main results as follows.
Theorem 1.1. Assume that (1.5) holds and 12<γ<1, and θ>0,Δn→0 and nΔ1+βn→∞ for some β>0 as n→∞. Then, we have, as n→∞,
and
Theorem 1.2. Assume that (1.5) holds and 12<γ<1, and θ>0,Δn→0 and nΔ1+βn→∞ for some β>0 as n→∞. Then, we have
(1) for any q≥0,
(2) if nΔ3n→0 as n→∞, then
Theorem 1.3. Assume that (1.5) holds and 12<γ<1, and θ>0,Δn→0 and nΔ1+βn→∞ for some β>0 as n→∞. Then, we have
(1) for any q≥0,
(2) if nΔ3n→0 as n→∞, then
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.
2.
Preliminaries and main lemmas
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
We know that the covariance of G can be written as
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
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
For every q≥1, 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)),h∈H,||h||H=1}, where Hq is the qth Hermite polynomial defined as Hq(x)=(−1)qex22dqdxq(e−x22). The mapping Iq(h⊗q)=Hq(G(h)) provides a linear isometry between the symmetric tensor product H⊙q (equipped with the modified norm ||⋅||H⊙q=√q!||⋅||H⊗q) and Hq. Specifically, for all f,g∈H⊙q and q≥1, one has
For the multiple stochastic integral Iq(f), it has the following property: for any p≥2,
where c(p,q) is a positive constant only depending on p and q.
It is easy to obtain the solution of (1.1):
where the integral with respect to Y(1)s is a Young integral (see Young (1936)).
Denote
then,
By (2.6) and dY(1)s=e−sdGas, as=γesγ and let as=u, we get
By (2.2) and 12<γ<1, we have, for 0≤s<t,
In order to prove Theorems 1.1-1.3, we need the following some lemmas.
Lemma 2.1. Let η={ηt,t≥0} 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 (γ−ϵ)-Hölder continuous paths, still denoted η in the sequel.
(2) As t→∞,
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}n∈N be a sequence of random variables. Suppose that, for every p≥1, there exists a constant cp>0 such that, for all n∈N,
Then, for all ϵ>0, there exists a random variable ξϵ such that, for any n∈N,
moreover, E[|ξϵ|p]<∞ for all p≥1.
Proof. See the proof in Kloeden and Neuenkirch (2007).
Lemma 2.3. Let
and
Then, we have
Assume that (1.5) holds and 12<γ<1, and θ>0,Δn→0 and nΔ1+βn→∞ for some β>0 as n→∞, then,
In particular, as n→∞, we have
where η∞ is defined by (2.10).
Proof. By (2.7), we obtain
which shows that (2.13) holds.
Using (2.9) and (1.5) and making the change of variables r=uati−1, s=vati−1, we get
By (2.10) and η is Gaussian, we also obtain, for every p≥1,
where c(p) is a constant depending on p. In fact, we have, by Cauchy-Schwarz inequality,
the last inequality comes from the fact: if ξ is Gaussian and p≥2, then
By (2.17) and (2.16) and the Minkowski inequality, we deduce that
where c is a constant, and the last inequality comes from the fact: as x→0,
Note that for any α>0, as n→∞,
In fact,
Since nΔ1+βn→∞ for some β>0 as n→∞, we have
In addition,
Thus (2.19) is obtained from (2.20)–(2.22).
By (2.19), we have, for any δ>0,
Hence (2.18) becomes
By Lemma 2.2, there exists a random variable ξα such that
for all n∈N. Moreover, E|ξα|p<∞ for all p≥1. Therefore (2.14) holds.
Finally, we can get (2.15) using (2.13), (2.14), (2.10) and limn→∞Δne2θΔn−1=12θ. The proof of Lemma 2.3 is finished.
3.
Proofs of Theorems
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,
where
and
Proof of Theorem 1.1. For (1.6), according to (2.15) and (3.1), it suffices to prove that, as n→∞,
By the Minkowski inequality and (2.16), we have
where c is a generic constant, and the last inequality also comes from the fact: as x→0,
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 limt→∞ηt=η∞ a.s., and this completes the proof of Theorem 1.1.
Proof of Theorem 1.2. (1) Firstly, we consider the case q≥12. By (3.1), we get
Since eθΔn−1−θΔn∼θ22Δ2n as Δn→0, we obtain
Because nΔ1+βn→∞ for some β>0 as n→∞ and
Then, we have
On the other hand, by (3.4), we get, as n→∞,
since γ>12. Consequently, by (3.5)–(3.7) and (2.15), for any q≥12, ΔqneθTn(ˆθn−θ) is not tight.
For the case 0≤q<12, note that
and limn→∞Δq−12n=∞ and the previous case q=12. The proof of (1.8) is finished.
(2) By (3.1), it is obvious that
Since nΔ3n→0 and eθΔn−1−θΔnΔ2n→θ22, we have, as n→∞,
Furthermore, by (3.4), we get, as n→∞,
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 q≥12, by (3.2) and (2.13), we obtain
By (2.16), we have, as n→∞,
We also get
since limn→∞nΔ1+βn=∞ for some β>0, limn→∞eθTnTq+1βn=∞, limn→∞e2θΔn−1−2θΔnΔ2n=2θ2 and limn→∞Δne2θΔn−1=12θ.
Moreover, by (2.18), we deduce, as n→∞,
where c(2) denotes c(p)=c(2) in (2.18). Combining (3.11)–(3.14) and (2.15), we conclude that for every q≥12, ΔqneθTn(˜θn−θ) is not tight.
For the case 0≤q<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
Similarly to (3.12)–(3.14), we obtain, as n→∞,
and
By (3.15)–(3.18), we can easily get (1.11).
Thus, we finish the proof of Theorem 1.3.
4.
Applications to fractional Gaussian processes
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.
4.1. Fractional Brownian motion
The fractional Brownian motion (fBm) BH={BHt,t≥0} with Hurst parameter H∈(0,1) is defined as a centered Gaussian process starting from zero with covariance
Note that, when H=12, B12 is a standard Brownian motion. By (4.1), we get
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).
4.2. Sub-fractional Brownian motion
The sub-fractional Brownian motion (sfBm) SH={SHt,t≥0} with Hurst parameter H∈(0,1) is defined as a centered Gaussian process starting from zero with covariance
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
When 12<H<1, by (4.4), we get
which shows that the assumption (1.5) holds for sfBm. Hence Theorems 1.1–1.3 hold for the sfBm SH(12<H<1).
4.3. Bifractional Brownian motion
The bifractional Brownian motion (bfBm) BH,K={BH,Kt,t≥0} with parameters H∈(0,1) and K∈(0,1] is defined as a centered Gaussian process starting from zero with covariance
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
Since K≤1, when 1<2HK<2, we have
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).
4.4. Sub-bifractional Brownian motion
El-Nouty and Journˊe (2013) introduced the process SH,K={SH,Kt,t≥0} with indices H∈(0,1) and K∈(0,1], named the sub-bifractional Brownian motion (sbfBm) and defined as follows:
where {BH,Kt,t∈R} 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)=(2K−22HK−1)t2HK. Note that since (2H−1)K−1<K−1≤0, it follows that 2HK−1<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,t≥0,
and
where
So SH,K has (HK−ϵ)-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
Since K≤1, when 1<2HK<2, we have
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).
5.
Numerical simulations
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
Let ti=iΔn,i=0,1,⋯,n,Xt0=0, then,
for i=1,2,⋯,n.
Let
where ξi(i=0,1,⋯,n) are standard normal random variables.
We know that
and
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,
and
Case 2: if G is sfBm SH(12<H<1), then (5.1) and (5.2) become (5.5) and (5.6), respectively,
and
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,
and
Case 4: if G is sbfBm SH,K(1<2HK<2), then (5.1) and (5.2) become (5.9) and (5.10), respectively,
and
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 1–4 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.
6.
Conclusions
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.
Acknowledgments
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.
Conflict of interest
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.