This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The Berry-Esséen Upper Bounds of Vasicek Model Estimators

Yong Chen School of Mathematics and Statistics, Jiangxi Normal University, Nanchang, 330022, Jiangxi, China [email protected]  and  Yumin Cheng School of Mathematics and Statistics, Jiangxi Normal University, Nanchang, 330022, Jiangxi, China [email protected]
Abstract.

The Berry-Esséen upper bounds of moment estimators and least squares estimators of the mean and drift coefficients in Vasicek models driven by general Gaussian processes are studied. When studying the parameter estimation problem of Ornstein-Uhlenbeck (OU) process driven by fractional Brownian motion, the commonly used methods are mainly given by Kim and Park, they show the upper bound of Kolmogorov distance between the distribution of the ratio of two double Wiener-Itô stochastic integrals and the Normal distribution. The main innovation in this paper is extending the above ratio process, that is to say, the numerator and denominator respectively contain triple Wiener-Itô stochastic integrals at most. As far as we know, the upper bounds between the distribution of above estimators and the Normal distribution are novel.

Key words and phrases:
Vasicek model, Malliavin calculus, Central limit theorem, Berry-Esséen upper bounds
2020 Mathematics Subject Classification:
60H07,60G15,60G22

1. Introduction

Vasicek model is a type of 1-dimensional stochastic processes, it is used in various fields, such as economy, finance, environment. It was originally used to describe short-term interest rate fluctuations influenced by single market factors. Proposed by O. Vasicek [1], it is the first stochastic process model to describe the “mean reversion” characteristic of short-term interest rates. In the financial field, it can also be used as a random investment model in Wu et al.[2] and Han et al.[3].

Definition 1.

Consider the Vasicek model driven by general Gaussian process, it satisfies the following Stochastic Differential Equation (SDE):

dVt=k(μVt)dt+σdGt,t[0,T],\mathop{}\!\mathrm{d}V_{t}=k(\mu-V_{t})\mathop{}\!\mathrm{d}t+\sigma\mathop{}\!\mathrm{d}G_{t},\;\;\;t\in[0,T], (1)

where k,T>0,V0=0k,T>0,\ V_{0}=0 and G={Gt}t0G=\{G_{t}\}_{t\geq 0} is a general one-dimensional centered Gaussian process that satisfies 1.

This paper mainly focuses on the convergence rate of estimators of coefficient k,μk,\mu. Without loss of generality, we assume σ=1\sigma=1, then Vasicek model can be represent by the following form:

Vt=μ(1ekt)+0tek(ts)dGs.V_{t}=\mu(1-\mathrm{e}^{-kt})+\int^{t}_{0}e^{-k(t-s)}\mathop{}\!\mathrm{d}G_{s}.

When the coefficients in the drift function is unknown, an important problem is to estimate the drift coefficients based on the observation. Based on the Brownian motion, Fergusson and Platen [4] present the maximum likelihood estimators of coefficients in Vasicek model. When the Vasicek model driven by the fractional Brownian motion, Xiao and Yu [5] consider the least squares estimators and their asymptotic behaviors. When k>0k>0, Hu and Nualart [6] study the moment estimation problem.

Since the Gaussian process GtG_{t} mainly determines the trajectory properties of Vasicek model. Therefore, following the assumptions in Chen and Zhou [7], we make the following Hypothesis about GtG_{t}.

Hypothesis 1 ([7] Hypothesis 1.1).

Let β(12,1)\beta\in(\frac{1}{2},1) and ts[0,)t\neq s\in[0,\infty), Covariance function R(t,s)=𝔼[GtGs]R(t,s)=\mathbb{E}[G_{t}G_{s}] of Gaussian process GtG_{t} satisfies the following condition:

2tsR(t,s)\displaystyle\frac{\partial^{2}}{\partial t\partial s}R(t,s) =Cβ|ts|2β2+Ψ(t,s),\displaystyle=C_{\beta}\left|t-s\right|^{2\beta-2}+\Psi(t,s), (2)

where

|Ψ(t,s)|\displaystyle\left|\Psi(t,s)\right| Cβ|ts|β1,\displaystyle\leq C_{\beta}^{\prime}\left|ts\right|^{\beta-1},

β,Cβ>0,Cβ0\beta,\ C_{\beta}>0,C_{\beta}^{\prime}\geq 0 are constants independent with TT. Besides, R(0,t)=0R(0,t)=0 for any t0t\geq 0.

Remark.

The covariance functions of Gaussian processes such as fractional Brownian motion, subfractional Brownian motion and double fractional Brownian motion satisfy the above Hypothesis [7, Examples 1.5-1.8].

Assuming that there is only one trajectory (Vt,t0)(V_{t},{t\geq 0}), we can construct the least squares estimators (LSE) and the moment estimators (ME) (See [8, 9, 5, 10] for more details).

Proposition 1 ([11] (4) and (5)).

The estimator of μ\mu is the continuous-time sample mean:

μ^=1T0TVtdt.\hat{\mu}=\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t. (3)

The second moment estimator of kk is given by

k^=[1T0TVt2dt(1T0TVtdt)2CβΓ(2β1)]12β.\hat{k}=\biggl{[}\frac{\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}}{C_{\beta}\Gamma(2\beta-1)}\biggr{]}^{-\frac{1}{2\beta}}. (4)

Following from Xiao and Yu [5], we present the LSE in Vasicek model.

Proposition 2 ([11] (7) and (8)).

The LSE is motivated by the argument of minimize a quadratic function L(k,μ)L(k,\mu) of kk and μ\mu:

L(k,μ)=0T(Vtk(μVt))2dt,L(k,\mu)=\int_{0}^{T}\bigl{(}\overset{\cdot}{V_{t}}-k(\mu-V_{t})\bigr{)}^{2}\mathop{}\!\mathrm{d}t,

Solving the equation, we can obtain the LSE of kk and μ\mu, denoted by k^LS\hat{k}_{LS} and μ^LS\hat{\mu}_{LS} respectively.

k^LS=VT0TVtdtT0TVtdVtT0TVt2dt(0TVtdt)2,\hat{k}_{LS}=\dfrac{V_{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t-T\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}}{T\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-(\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}}, (5)
μ^LS=VT0TVt2dt0TVtdVt0TVtdtVT0TVtdtT0TVtdVt,\hat{\mu}_{LS}=\frac{V_{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t}{V_{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t-T\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}}, (6)

where the integral 0TVtdVt\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t} is an Itô-Skorohod integral.

Pei et al.[11] prove the following consistencies and central limit theorems (CLT) of estimators.

Theorem 3 ([11], Theorem 1.2).

When 1 is satisfied, both ME and LSE of μ,k\mu,k are strongly consistent, that is

limTμ^=μ,limTk^=k,a.s.;\displaystyle\lim_{T\to\infty}\hat{\mu}=\mu,\;\;\;\;\;\;\lim_{T\to\infty}\hat{k}=k,\;\;\;\mathrm{a.s.};
limTμ^LS=μ,limTk^LS=k,a.s..\displaystyle\lim_{T\to\infty}\hat{\mu}_{LS}=\mu,\;\;\;\lim_{T\to\infty}\hat{k}_{LS}=k,\;\;\;\mathrm{a.s.}.
Theorem 4 ([11], Theorem 1.3).

Assume 1 is satisfied. When GtG_{t} is self-similar and 𝔼[G12]=1\mathbb{E}[G_{1}^{2}]=1, μ^\hat{\mu} and μ^LS\hat{\mu}_{LS} are asymptotically normal as TT\to\infty, that is,

T1β(μ^μ)law𝒩(0,1/k2),T(μ^LSμ)law𝒩(0,1/k2).T^{1-\beta}(\hat{\mu}-\mu)\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,1/k^{2}),\;\;\;\sqrt{T}(\hat{\mu}_{LS}-\mu)\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,1/k^{2}).

When β(12,34)\beta\in(\frac{1}{2},\frac{3}{4}),

T(k^k)law𝒩(0,kσβ2/4β2),\sqrt{T}(\hat{k}-k)\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,k\sigma^{2}_{\beta}/4\beta^{2}),

where

σβ2=(4β1)[1+Γ(34β)Γ(4β1)Γ(2β)Γ(22β)].\displaystyle\sigma_{\beta}^{2}=(4\beta-1)\biggl{[}1+\frac{\Gamma(3-4\beta)\Gamma(4\beta-1)}{\Gamma(2\beta)\Gamma(2-2\beta)}\biggr{]}. (7)

Simalarly, T(k^LSk)\sqrt{T}(\hat{k}_{LS}-k) is also asymptotically normal as TT\to\infty:

T(k^LSk)\displaystyle\sqrt{T}(\hat{k}_{LS}-k) law𝒩(0,kσβ2).\displaystyle\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,k\sigma_{\beta}^{2}).

We now present the main Theorems for the whole paper, and their details are given in the following sections.

Theorem 5.

Let ZZ be a standard Normal random variable, and σβ2\sigma_{\beta}^{2} be the constant defined by (7). Assume β(1/2,3/4)\beta\in(1/2,3/4) and 1 is satisfied. When TT is large enough, there exists a constant Cβ,VC_{\beta,V} such that

supz\displaystyle\sup_{z\in\mathbb{R}} |(4β2Tkσβ2(k^k)z)(Zz)|Cβ,VTm,\displaystyle\left|\mathbb{P}\biggl{(}\sqrt{\frac{4\beta^{2}T}{k\sigma^{2}_{\beta}}}(\hat{k}-k)\leq z\biggr{)}-\mathbb{P}(Z\leq z)\right|\leq\frac{C_{\beta,V}}{T^{m}}, (8)
supz\displaystyle\sup_{z\in\mathbb{R}} |(Tkσβ2(k^LSk))(Zz)|Cβ,VT3/4β,\displaystyle\left|\mathbb{P}\biggl{(}\sqrt{\frac{T}{k\sigma_{\beta}^{2}}}(\hat{k}_{LS}-k)\biggr{)}-\mathbb{P}(Z\leq z)\right|\leq\frac{C_{\beta,V}}{T^{3/4-\beta}}, (9)

where m=min{1/3,(34β)/2}m=\min\{1/3,(3-4\beta)/2\}.

Next, we show the convergence speed of mean coefficient estimators μ^\hat{\mu} and μ^LS\hat{\mu}_{LS}.

Theorem 6.

Assume β(1/2,1)\beta\in(1/2,1), and GtG_{t} is a self-similar Gaussian process satisfying 1 and 𝔼[G12]=1\mathbb{E}[G_{1}^{2}]=1. Then there exists a constant Cβ,VC_{\beta,V} such that

supz\displaystyle\sup_{z\in\mathbb{R}} |(kTβ1(μ^μ)z)(Zz)|Cβ,VTβ/2,\displaystyle\left|\mathbb{P}\biggl{(}\frac{k}{T^{\beta-1}}(\hat{\mu}-\mu)\leq z\biggr{)}-\mathbb{P}(Z\leq z)\right|\leq\frac{C_{\beta,V}}{T^{\beta/2}}, (10)
supz\displaystyle\sup_{z\in\mathbb{R}} |(kTβ1(μ^LSμ))(Zz)|Cβ,VT(1β)/2.\displaystyle\left|\mathbb{P}\biggl{(}\frac{k}{T^{\beta-1}}(\hat{\mu}_{LS}-\mu)\biggr{)}-\mathbb{P}(Z\leq z)\right|\leq\frac{C_{\beta,V}}{T^{(1-\beta)/2}}. (11)

2. Preliminary

In this section, we recall some basic facts about Malliavin calculus with respect to Gaussian process. The reader is referred to [12, 13, 14] for a more detailed explanation. Let G={Gt,t[0,T]}G=\{G_{t},t\in[0,T]\} be a continuous centered Gaussian process with G0=0G_{0}=0 and covariance function

𝔼(GtGs)=R(s,t),s,t[0,T],\mathbb{E}(G_{t}G_{s})=R(s,t),\,\,s,t\in[0,T], (12)

defined on a complete probability space (Ω,,)(\Omega,\mathscr{F},\mathbb{P}), where \mathscr{F} is generated by the Gaussian family GG. Denote \mathcal{E} as the the space of all real valued step functions on [0,T][0,T]. The Hilbert space \mathfrak{H} is defined as the closure of \mathcal{E} endowed with the inner product:

𝟙[a,b),𝟙[c,d)=𝔼((GbGa)(GdGc)).\langle\mathbbm{1}_{[a,b)},\mathbbm{1}_{[c,d)}\rangle_{\mathfrak{H}}=\mathbb{E}((G_{b}-G_{a})(G_{d}-G_{c})). (13)

We denote G={G(h),h}G=\{G(h),h\in{\mathfrak{H}}\} as the isonormal Gaussian process on the probability space, indexed by the elements in \mathfrak{H}, which satisfies the following isometry relationship:

𝔼(G(h))=0,𝔼[G(g)G(h)]=g,h,g,h.\mathbb{E}\bigl{(}G(h)\bigr{)}=0,\,\,\mathbb{E}[G(g)G(h)]=\langle g,h\rangle_{\mathfrak{H}},\quad\forall\ g,h\in\mathfrak{H}. (14)

The following Proposition shows the inner products representation of the Hilbert space \mathfrak{H} [15].

Proposition 7 ([7] Proposition 2.1).

Denote 𝒱[0,T]\mathcal{V}_{[0,T]} as the set of bounded variation functions on [0,T][0,T]. Then 𝒱[0,T]\mathcal{V}_{[0,T]} is dense in \mathfrak{H} and

f,g=[0,T]2R(t,s)vf(dt)vg(ds),f,g𝒱[0,T],\langle f,g\rangle_{\mathfrak{H}}=\int_{[0,T]^{2}}R(t,s){v}_{f}(\mathop{}\!\mathrm{d}t){v}_{g}(\mathop{}\!\mathrm{d}s),\quad\forall\ f,g\in\mathcal{V}_{[0,T]},

where vgv_{g} is the Lebesgue-Stieljes signed measure associated with g0g^{0} defined as

g0={g(x),ifx[0,T];0,otherwise.g^{0}=\left\{\begin{array}[]{rcr}g(x),&&{\text{if}\ x\in[0,T];}\\ 0,&&\text{otherwise}.\\ \end{array}\right.

When the covariance function R(t,s)R(t,s) satisfies 1,

f,g=[0,T]2f(t)g(s)2R(t,s)tsdtds,f,g𝒱[0,T].\displaystyle\langle f,g\rangle_{\mathfrak{H}}=\int_{[0,T]^{2}}f(t)g(s)\frac{\partial^{2}R(t,s)}{\partial t\partial s}\mathop{}\!\mathrm{d}t\mathop{}\!\mathrm{d}s,\quad\forall\ f,g\in\mathcal{V}_{[0,T]}. (15)

Furthermore, the norm \left\|\cdot\right\|_{\mathfrak{H}} of the elements in \mathfrak{H} can be induced naturally:

ϕ2=[0,T]2ϕ(r1)ϕ(r2)2R(r1,r2)r1r2dr1dr2,ϕ.\left\|\phi\right\|_{\mathfrak{H}}^{2}=\int_{[0,T]^{2}}\phi(r_{1})\phi(r_{2})\frac{\partial^{2}R(r_{1},r_{2})}{\partial r_{1}\partial r_{2}}\mathop{}\!\mathrm{d}r_{1}\mathop{}\!\mathrm{d}r_{2},\;\;\;\forall\ \phi\in\mathfrak{H}.
Remark ([7] Notation 1).

Let CβC_{\beta} and CβC_{\beta}^{\prime} be the constants given in 1. For any ϕ(r)𝒱[0,T]\phi(r)\in\mathcal{V}_{[0,T]}, we define two norms as

ϕ12\displaystyle\left\|\phi\right\|_{\mathfrak{H}_{1}}^{2} =Cβ[0,T]2ϕ(r1)ϕ(r2)|r1r2|2β2dr1dr2,\displaystyle=C_{\beta}\int_{[0,T]^{2}}\phi(r_{1})\phi(r_{2})\left|r_{1}-r_{2}\right|^{2\beta-2}\mathop{}\!\mathrm{d}r_{1}\mathop{}\!\mathrm{d}r_{2},
ϕ22\displaystyle\left\|\phi\right\|_{\mathfrak{H}_{2}}^{2} =Cβ[0,T]2|ϕ(r1)ϕ(r2)|(r1r2)β1dr1dr2.\displaystyle=C_{\beta}^{\prime}\int_{[0,T]^{2}}\left|\phi(r_{1})\phi(r_{2})\right|(r_{1}r_{2})^{\beta-1}\mathop{}\!\mathrm{d}r_{1}\mathop{}\!\mathrm{d}r_{2}.

For any φ(r,s)\varphi(r,s) in [0,T]2[0,T]^{2}, define an operator KK from 𝒱[0,T]2\mathcal{V}_{[0,T]}^{\otimes 2} to 𝒱[0,T]\mathcal{V}_{[0,T]} to be

(Kφ)(r)=0T|φ(r,u)|uβ1du.(K\varphi)(r)=\int_{0}^{T}\left|\varphi(r,u)\right|u^{\beta-1}\mathop{}\!\mathrm{d}u. (16)
Proposition 8 ([7] Proposition 3.2).

Suppose that 1 holds, then for any ϕ(r)𝒱[0,T]\phi(r)\in\mathcal{V}_{[0,T]},

|ϕ2ϕ12|ϕ22,\left|\left\|\phi\right\|_{\mathfrak{H}}^{2}-\left\|\phi\right\|_{\mathfrak{H}_{1}}^{2}\right|\leq\left\|\phi\right\|_{\mathfrak{H}_{2}}^{2}, (17)

and for any φ,ψ(𝒱[0,T])2\varphi,\psi\in(\mathcal{V}_{[0,T]})^{\odot 2},

|ϕ22ϕ122|\displaystyle\left|\left\|\phi\right\|_{\mathfrak{H}^{\otimes{2}}}^{2}-\left\|\phi\right\|_{\mathfrak{H}_{1}^{\otimes{2}}}^{2}\right| ϕ222+2CβKφ12,\displaystyle\leq\left\|\phi\right\|_{\mathfrak{H}_{2}^{\otimes{2}}}^{2}+2C_{\beta}^{\prime}\left\|K\varphi\right\|_{\mathfrak{H}_{1}}^{2},
|φ,ψ2φ,ψ12|\displaystyle\left|\langle\varphi,\psi\rangle_{\mathfrak{H}^{\otimes{2}}}-\langle\varphi,\psi\rangle_{\mathfrak{H}_{1}^{\otimes{2}}}\right| |φ,ψ22|+2Cβ|Kφ,Kψ1|.\displaystyle\leq\left|\langle\varphi,\psi\rangle_{\mathfrak{H}_{2}^{\otimes{2}}}\right|+2C_{\beta}^{\prime}\left|\langle K\varphi,K\psi\rangle_{\mathfrak{H}_{1}}\right|.

Let p\mathfrak{H}^{\otimes p} and p\mathfrak{H}^{\odot p} be the pp-th tensor product and the pp-th symmetric tensor product of \mathfrak{H}. For every p1p\geq 1, denote p\mathcal{H}_{p} as the pp-th Wiener chaos of GG. It is defined as the closed linear subspace of L2(Ω)L^{2}(\Omega) generated by {Hp(G(h)):h,h=1}\{H_{p}(G(h)):h\in{\mathfrak{H}},\left\|h\right\|_{\mathfrak{H}}=1\}, where HpH_{p} is the pp-th Hermite polynomial. Let hh\in\mathfrak{H} such that h=1\left\|h\right\|_{\mathfrak{H}}=1, then for every p1p\geq 1 and hh\in\mathfrak{H},

Ip(hp)=Hp(G(h)),I_{p}(h^{\otimes p})=H_{p}\bigl{(}G(h)\bigr{)},

where Ip()I_{p}(\cdot) is the pp-th Wiener-Itô stochastic integral.

Denote {ek,k1}\{e_{k},k\geq 1\} as a complete orthonormal system in \mathfrak{H}. The qq-th contraction between fmf\in\mathfrak{H}^{\odot m} and gng\in\mathfrak{H}^{\odot n} is an element in (m+n2q)\mathfrak{H}^{\otimes(m+n-2q)}:

fqg=i1,,iq=1f,ei1,,eiqqg,ei1,,eiqq,q=1,,mn.\displaystyle f\otimes_{q}g=\sum_{i_{1},\cdots,i_{q}=1}^{\infty}\langle f,e_{i_{1}},\cdots,e_{i_{q}}\rangle_{\mathfrak{H}^{\otimes q}}\otimes\langle g,e_{i_{1}},\cdots,e_{i_{q}}\rangle_{\mathfrak{H}^{\otimes q}},\;\;\;q=1,\cdots,m\wedge n.

The following proposition shows the product formula for the multiple integrals.

Proposition 9 ([12] Theorem 2.7.10).

Let fpf\in\mathfrak{H}^{\odot p} and gqg\in\mathfrak{H}^{\odot q} be two symmetric function. Then

Ip(f)Iq(g)=r=0pqr!(pr)(qr)Ip+q2r(f~rg),I_{p}(f)I_{q}(g)=\sum_{r=0}^{p\wedge q}r!\tbinom{p}{r}\tbinom{q}{r}I_{p+q-2r}(f\tilde{\otimes}_{r}g), (18)

where f~rgf\tilde{\otimes}_{r}g is the symmetrization of frgf{\otimes}_{r}g.

We then introduce the derivative operator and the divergence operator. For these details, see sections 2.3-2.5 of [12]. Let 𝒯\mathscr{T} be the class of smooth random variables of the form:

F=f(G(ψ1),G(ψ2),,G(ψn)),F=f\bigl{(}G(\psi_{1}),G(\psi_{2}),\cdots,G(\psi_{n})\bigr{)},

where n1n\geq 1, f𝒞b(n)f\in\mathcal{C}_{b}^{\infty}(\mathbb{R}^{n}) which partial derivatives have at most polynomial growth, and for i=1,,ni=1,\cdots,n, ψi\psi_{i}\in\mathfrak{H}. Then, the Malliavin derivative of FF (with respect to GG) is the element of L2(Ω,)L^{2}(\Omega,\mathfrak{H}) defined by

DF=i=1nfxi(G(ψ1),,G(ψn))ψi.DF=\sum_{i_{=}1}^{n}\frac{\partial f}{\partial x_{i}}\bigl{(}G(\psi_{1}),\cdots,G(\psi_{n})\bigr{)}\psi_{i}.

Given q[1,)q\in[1,\infty) and integer p1p\geq 1, let 𝔻p,q\mathbb{D}^{p,q} denote the closure of 𝒯\mathscr{T} with respect to the norm

F𝔻p,q=[𝔼(|F|q)+k=1p𝔼(DkFkq)]1/q.\left\|F\right\|_{\mathbb{D}^{p,q}}=\Bigl{[}\mathbb{E}(\left|F\right|^{q})+\sum_{k=1}^{p}\mathbb{E}(\left\|D^{k}F\right\|^{q}_{\mathfrak{H}\otimes k})\Bigr{]}^{1/q}.

Denote δ\delta (the divergence operator) as the adjoint of DD. The domain of δ\delta is composed of those elements:

|𝔼[DpF,up]|C[𝔼(|F|2)]12,F𝔻p,2,\left|\mathbb{E}[\langle D^{p}F,u\rangle_{\mathfrak{H}\otimes p}]\right|\leq C[\mathbb{E}(\left|F\right|^{2})]^{\frac{1}{2}},\;\;\;\forall\ F\in\mathbb{D}^{p,2},

and is denoted by Dom(δ)\mathrm{Dom}(\delta). If uDom(δ)u\in\mathrm{Dom}(\delta), then δ(u)\delta(u) is the unique element of L2(Ω)L^{2}(\Omega) characterized by the duality formula:

𝔼[Fδ(u)]=𝔼(DF,u),F𝔻1,2.\mathbb{E}[F\delta(u)]=\mathbb{E}(\langle DF,u\rangle_{\mathfrak{H}}),\;\;\;\forall\ F\in\mathbb{D}^{1,2}.

We now introduce the infinitesimal generator LL of the Ornstein-Uhlenbeck semigroup. Let FL2(Ω)F\in L^{2}(\Omega) be a square integrable random variable. Denote 𝕁n:L2(Ω)n\mathbb{J}_{n}:L^{2}(\Omega)\to\mathcal{H}_{n} as the orthogonal projection on the nn-th Wiener chaos n\mathcal{H}_{n}. The operator LL is defined by LF=n=0n𝕁nFLF=-\sum_{n=0}^{\infty}n\mathbb{J}_{n}F. The domain of LL is

Dom(L):={FL2(Ω),p=1p2𝔼[𝕁p(F)2]}.\displaystyle\mathrm{Dom}(L):=\{F\in L^{2}(\Omega),\sum_{p=1}^{\infty}p^{2}\mathbb{E}[\mathbb{J}_{p}(F)^{2}]\leq\infty\}.

For any FL2(Ω)F\in L^{2}(\Omega), define L1F=n=11n𝕁nFL^{-1}F=-\sum_{n=1}^{\infty}\frac{1}{n}\mathbb{J}_{n}F. L1L^{-1} is called the Pseudo-inverse of LL. Note that L1FDom(L)L^{-1}F\in\mathrm{Dom}(L) and LL1F=F𝔼(F)LL^{-1}F=F-\mathbb{E}(F) holds for any FL2(Ω)F\in L^{2}(\Omega).

The following Lemma 10 provides the Berry-Esséen upper bound on the sum of two random variables.

Lemma 10 ([16] Lemma 2).

For any variable ξ,η\xi,\eta and aa\in\mathbb{R}, the following inequality holds:

supz|(ξ+ηz)Φ(z)|supz|(ξz)Φ(z)|+(|η|>a)+a2π,\sup_{z\in\mathbb{R}}\left|\mathbb{P}(\xi+\eta\leq z)-\Phi(z)\right|\leq\sup_{z\in\mathbb{R}}\left|\mathbb{P}(\xi\leq z)-\Phi(z)\right|+\mathbb{P}(\left|\eta\right|>a)+\frac{a}{\sqrt{2\pi}}, (19)

where Φ(z)\Phi(z) is the standard Normal distribution function.

Using Malliavin calculus, Kim and Park [17] provide the Berry-Esséen upper bound of the quotient of two random variables.

Let FT𝔻1,2F_{T}\in\mathbb{D}^{1,2} be a zero-mean process, and GT𝔻1,2G_{T}\in\mathbb{D}^{1,2} satisfies GT>0G_{T}>0 a.s.. For simplicity, we define the following four functions:

Ψ1(T)\displaystyle\Psi_{1}(T) =1(𝔼GT)2𝔼[((𝔼GT)2(DFT,DL1FT))2],\displaystyle=\frac{1}{(\mathbb{E}{G_{T}})^{2}}\sqrt{\mathbb{E}\Bigl{[}\Bigl{(}(\mathbb{E}G_{T})^{2}-(\langle DF_{T},-DL^{-1}F_{T}\rangle_{\mathfrak{H}})\Bigr{)}^{2}\Bigr{]}},
Ψ2(T)\displaystyle\Psi_{2}(T) =1(𝔼GT)2𝔼[DFT,DL1(GT𝔼GT)2],\displaystyle=\frac{1}{(\mathbb{E}{G_{T}})^{2}}\sqrt{\mathbb{E}\Bigl{[}\langle DF_{T},-DL^{-1}(G_{T}-\mathbb{E}G_{T})\rangle_{\mathfrak{H}}^{2}\Bigr{]}},
Ψ3(T)\displaystyle\Psi_{3}(T) =1(𝔼GT)2𝔼[DGT,DL1FT2],\displaystyle=\frac{1}{(\mathbb{E}{G_{T}})^{2}}\sqrt{\mathbb{E}\Bigl{[}\langle DG_{T},-DL^{-1}F_{T}\rangle_{\mathfrak{H}}^{2}\Bigr{]}},
Ψ4(T)\displaystyle\Psi_{4}(T) =1(𝔼GT)2𝔼[DGT,DL1(GT𝔼GT)2].\displaystyle=\frac{1}{(\mathbb{E}{G_{T}})^{2}}\sqrt{\mathbb{E}\Bigl{[}\langle DG_{T},-DL^{-1}(G_{T}-\mathbb{E}G_{T})\rangle_{\mathfrak{H}}^{2}\Bigr{]}}.
Theorem 11 ([17] Theorem 2 and Corollary 1).

Let ZZ be a standard Normal variable. Assuming that for every zz\in\mathbb{R}, FT+zGTF_{T}+zG_{T} has an absolutely continuous law with respect to Lebesgue measure and Ψi(T)0\Psi_{i}(T)\to 0, i=1,,4i=1,\cdots,4, as TT\to\infty. Then, there exists a constant cc such that for TT large enough,

supz|(FTGTz)(Zz)|\displaystyle\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\frac{F_{T}}{G_{T}}\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right| cmaxi=1,,4Ψi(T).\displaystyle\leq c\cdot\max_{i=1,\cdots,4}\Psi_{i}(T).

3. Berry-Esséen upper bounds of moment estimators

In this section, we will prove the Berry-Esséen upper bounds of Vasicek model moment estimators μ^\hat{\mu} and k^\hat{k}. For the convenience of the following discussion, we first define A(z)A(z):

A(z):=(kTβ1(μ^μ)z)(Zz),\displaystyle A(z):=\mathbb{P}\biggl{(}\frac{k}{T^{\beta-1}}(\hat{\mu}-\mu)\leq z\biggr{)}-\mathbb{P}(Z\leq z), (20)

where Z𝒩(0,1)Z\sim\mathcal{N}(0,1) is a standard Normal variable. Next, we introduce the CLT of μ^\hat{\mu}.

Theorem 12 ([11] Proposition 4.19).

Assume β(1/2,1)\beta\in(1/2,1), and GtG_{t} is a self-similar Gaussian process satisfying 1 and 𝔼[G12]=1\mathbb{E}[G_{1}^{2}]=1. Then T1β(μ^μ)T^{1-\beta}(\hat{\mu}-\mu) is asymptotically normal as TT\to\infty:

T1β(μ^μ)=μkekT1Tβ+1kGTZTTβlaw𝒩(0,1/k2),\displaystyle T^{1-\beta}(\hat{\mu}-\mu)=\frac{\mu}{k}\frac{\mathrm{e}^{-kT}-1}{T^{\beta}}+\frac{1}{k}\frac{G_{T}-Z_{T}}{T^{\beta}}\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,1/k^{2}), (21)

where

ZT=I1(ek(Ts)𝟙[0,T](s))Z_{T}=I_{1}\bigl{(}\mathrm{e}^{-k(T-s)}\mathbbm{1}_{[0,T]}(s)\bigr{)}

is stochastic integral with respect to GtG_{t}.

Following from the above Theorem, we can obtain the expanded form of (20):

A(z)\displaystyle A(z) =(kT1β(μ^μ)z)(Zz)\displaystyle=\mathbb{P}\Bigl{(}kT^{1-\beta}(\hat{\mu}-\mu)\leq z\Bigr{)}-\mathbb{P}(Z\leq z)
=(GTZT+μ(ekT1)Tβz)(Zz).\displaystyle=\mathbb{P}\biggl{(}\frac{G_{T}-Z_{T}+\mu(\mathrm{e}^{-kT}-1)}{T^{\beta}}\leq z\biggr{)}-\mathbb{P}(Z\leq z).

Then, we can prove the convergence speed of μ^\hat{\mu}.

Proof of formula (10).

Let a=Tβ/2a=T^{-\beta/2}, According to Lemma 10, we have

supz|A(z)|\displaystyle\sup_{z\in\mathbb{R}}\left|A(z)\right| supz|(GTTβz)(Zz)|\displaystyle\leq\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\frac{G_{T}}{T^{\beta}}\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right|
+(|ZT+μ(ekT1)Tβ|>Tβ2)+Tβ22π.\displaystyle\ \ \ +\mathbb{P}\biggl{(}\left|\frac{-Z_{T}+\mu(\mathrm{e}^{-kT}-1)}{T^{\beta}}\right|>T^{-\frac{\beta}{2}}\biggr{)}+\frac{T^{-\frac{\beta}{2}}}{\sqrt{2\pi}}.

Since GTG_{T} is self-similar, (GT/Tβ)(G_{T}/{T^{\beta}}) is standard Normal variable,

supz|(GTTβz)(Zz)|=0.\displaystyle\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\frac{G_{T}}{T^{\beta}}\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right|=0.

Following from Chebyshev inequality, we can obtain

(|ZT+μ(ekT1)Tβ|>Tβ2)=(|ZTμ(ekT1)|>Tβ2)C1Tβ,\displaystyle\mathbb{P}\biggl{(}\left|\frac{-Z_{T}+\mu(\mathrm{e}^{-kT}-1)}{T^{\beta}}\right|>T^{-\frac{\beta}{2}}\biggr{)}=\mathbb{P}\bigl{(}\left|Z_{T}-\mu(\mathrm{e}^{-kT}-1)\right|>T^{\frac{\beta}{2}}\bigr{)}\leq\frac{C_{1}}{T^{\beta}},

where

C1=𝔼(|ZTμ(ekT1)|2).\displaystyle C_{1}=\mathbb{E}\bigl{(}\left|Z_{T}-\mu(\mathrm{e}^{-kT}-1)\right|^{2}\bigr{)}.

The Proposition 3.10 of [11] ensures that C1C_{1} is bounded. Combining the above results, we have

supz|A(z)|12πTβ/2+C1Tβ.\displaystyle\sup_{z\in\mathbb{R}}\left|A(z)\right|\leq\frac{1}{\sqrt{2\pi}T^{\beta/2}}+\frac{C_{1}}{T^{\beta}}. (22)

When TT is sufficiently large, there exists the constant Cβ,VC_{\beta,V} such that the formula (10) holds. ∎

Similarly, we review the central limit theorem of k^\hat{k}.

Theorem 13 ([11] Proposition 4.18).

Assume β(1/2,3/4)\beta\in(1/2,3/4) and GtG_{t} is a Gaussian process satisfying 1. Then T(k^k)\sqrt{T}(\hat{k}-k) is asymptotically normal as TT\to\infty:

T(k^k)=T([1T0TVt2𝑑t(1T0TVt𝑑t)2CβΓ(2β1)]12βk)law𝒩(0,kσβ2/4β2).\displaystyle\sqrt{T}(\hat{k}-k)=\sqrt{T}\left(\left[\frac{\frac{1}{T}\int_{0}^{T}V_{t}^{2}dt-(\frac{1}{T}\int_{0}^{T}V_{t}dt)^{2}}{C_{\beta}\Gamma(2\beta-1)}\right]^{-\frac{1}{2\beta}}-k\right)\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,k\sigma^{2}_{\beta}/4\beta^{2}).

The following Lemma shows the upper bound of the expectation of (0TeksdGs)2(\int_{0}^{T}\mathrm{e}^{-ks}\mathop{}\!\mathrm{d}G_{s})^{2}.

Lemma 14.

Let MTM_{T} be the process defined by

MT=I1(eku𝟙[0,T](u))=0TeksdGs.M_{T}=I_{1}\bigl{(}\mathrm{e}^{-ku}\mathbbm{1}_{[0,T]}(u)\bigr{)}=\int_{0}^{T}\mathrm{e}^{-ks}\mathop{}\!\mathrm{d}G_{s}.

When β(1/2,1)\beta\in(1/2,1), there exists constant CC independent of TT such that

𝔼(MT2)C.\mathbb{E}(M^{2}_{T})\leq C. (23)
Proof.

According to (14) and (17), we can obtain

𝔼[|MT|2]=mT2mT12+mT22,\displaystyle\mathbb{E}[\left|M_{T}\right|^{2}]=\left\|m_{T}\right\|_{\mathfrak{H}}^{2}\leq\left\|m_{T}\right\|^{2}_{\mathfrak{H}_{1}}+\left\|m_{T}\right\|^{2}_{\mathfrak{H}_{2}},

where mT(u)=eku𝟙[0,T](u)m_{T}(u)=\mathrm{e}^{-ku}\mathbbm{1}_{[0,T]}(u).

It is easy to see that

mT12\displaystyle\left\|m_{T}\right\|^{2}_{\mathfrak{H}_{1}} =2Cβ0uvTek(u+v)|uv|2β2dudv\displaystyle=2C_{\beta}\int_{0\leq u\leq v\leq T}\mathrm{e}^{-k(u+v)}\left|u-v\right|^{2\beta-2}\mathop{}\!\mathrm{d}u\mathop{}\!\mathrm{d}v
C0T(0Tek(2x+y)x2β2dx)dyC,\displaystyle\leq C\int_{0}^{T}(\int_{0}^{T}\mathrm{e}^{-k(2x+y)}x^{2\beta-2}\mathop{}\!\mathrm{d}x)\mathop{}\!\mathrm{d}y\leq C^{\prime}, (24)

where CC^{\prime} is a constant. Also, we have

mT22\displaystyle\left\|m_{T}\right\|^{2}_{\mathfrak{H}_{2}} =Cβ0Tekuuβ1du0Tekvvβ1dvC′′.\displaystyle=C^{\prime}_{\beta}\int_{0}^{T}\mathrm{e}^{-ku}u^{\beta-1}\mathop{}\!\mathrm{d}u\int_{0}^{T}\mathrm{e}^{-kv}v^{\beta-1}\mathop{}\!\mathrm{d}v\leq C^{\prime\prime}. (25)

Combining the above two formulas, we obtain (23). ∎

Denote B(z)B(z) as

B(z):\displaystyle B(z): =(4β2Tkσβ2(k^k)z)(Zz).\displaystyle=\mathbb{P}\biggl{(}\sqrt{\frac{4\beta^{2}T}{k\sigma^{2}_{\beta}}}(\hat{k}-k)\leq z\biggr{)}-\mathbb{P}(Z\leq z).

Then we can obtain the Berry-Esséen upper bound of ME k^\hat{k}.

Proof of formula (8).

According to [11] Proposition 4.18, we have

B(z)\displaystyle B(z) =(4β2Tkσβ2(k^k)z)(Zz)\displaystyle=\mathbb{P}\biggl{(}\sqrt{\frac{4\beta^{2}T}{k\sigma^{2}_{\beta}}}(\hat{k}-k)\leq z\biggr{)}-\mathbb{P}(Z\leq z)
=(k^kkσβ24β2Tz)(Zz)\displaystyle=\mathbb{P}\biggl{(}\hat{k}-k\leq\sqrt{\frac{k\sigma^{2}_{\beta}}{4\beta^{2}T}}z\biggr{)}-\mathbb{P}(Z\leq z)
=(1T0TVt2dt(1T0TVtdt)2CβΓ(2β1)(kσβ24β2Tz+k)2β)(Zz)\displaystyle=\mathbb{P}\biggl{(}\frac{\frac{1}{T}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-\bigl{(}\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t\bigr{)}^{2}}{C_{\beta}\Gamma(2\beta-1)}\geq\biggl{(}\sqrt{\frac{k\sigma^{2}_{\beta}}{4\beta^{2}T}}z+k\biggr{)}^{-2\beta}\biggr{)}-\mathbb{P}(Z\leq z)
=(1T(0TVt2dt(1T0TVtdt)2)α\displaystyle=\mathbb{P}\biggl{(}\frac{1}{T}\bigl{(}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}\bigr{)}-\alpha
CβΓ(2β1)[(kσβ24β2Tz+k)2βk2β])(Zz)\displaystyle\;\;\;\;\;\;\;\;\;\geq C_{\beta}\Gamma(2\beta-1)\biggl{[}\biggl{(}\sqrt{\frac{k\sigma^{2}_{\beta}}{4\beta^{2}T}}z+k\biggr{)}^{-2\beta}-k^{-2\beta}\biggr{]}\biggr{)}-\mathbb{P}(Z\leq z)
=(1T(0TVt2dt(1T0TVtdt)2)αα[(1+zσβ2βkT)2β1])\displaystyle=\mathbb{P}\biggl{(}\frac{1}{T}\bigl{(}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}\bigr{)}-\alpha\geq\alpha\biggl{[}\biggl{(}1+\frac{z\sigma_{\beta}}{2\beta\sqrt{kT}}\biggr{)}^{-2\beta}-1\biggr{]}\biggr{)}
(Zz).\displaystyle\ \ \ -\mathbb{P}(Z\leq z).

We denote Φ¯(z)\overline{\Phi}(z) as the tail probability 1(Zz)1-\mathbb{P}(Z\leq z) and

ν=kTσβ2[(1+zσβ2βkT)2β1].\nu=\sqrt{\frac{kT}{\sigma_{\beta}^{2}}}\biggl{[}\biggl{(}1+\frac{z\sigma_{\beta}}{2\beta\sqrt{kT}}\biggr{)}^{-2\beta}-1\biggr{]}.

Then we can obtain

|B(z)|\displaystyle\left|B(z)\right| =|(kTα2σβ2[1T(0TVt2dt(1T0TVtdt)2)α]ν)(Zz)|\displaystyle=\left|\mathbb{P}\biggl{(}\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}\biggl{[}\frac{1}{T}\bigl{(}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}\bigr{)}-\alpha\biggr{]}\geq\nu\biggr{)}-\mathbb{P}(Z\leq z)\right|
|(kTα2σβ2[1T(0TVt2dt(1T0TVtdt)2)α]ν)Φ¯(ν)|\displaystyle\leq\left|\mathbb{P}\biggl{(}\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}\biggl{[}\frac{1}{T}\bigl{(}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}\bigr{)}-\alpha\biggr{]}\geq\nu\biggr{)}-\overline{\Phi}(\nu)\right|
+|Φ¯(ν)(Zz)|\displaystyle\ \ \ +\left|\overline{\Phi}(\nu)-\mathbb{P}(Z\leq z)\right|
=|D(ν)|+|Φ¯(ν)(Zz)|.\displaystyle=\left|D(\nu)\right|+\left|\overline{\Phi}(\nu)-\mathbb{P}(Z\leq z)\right|. (26)

Denote D(v)D(v) as

D(ν):\displaystyle D(\nu): =|(kTα2σβ2[1T(0TVt2dt(1T0TVtdt)2)α]ν)Φ¯(ν)|,\displaystyle=\left|\mathbb{P}\biggl{(}\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}\biggl{[}\frac{1}{T}\bigl{(}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}\bigr{)}-\alpha\biggr{]}\geq\nu\biggr{)}-\overline{\Phi}(\nu)\right|,

where ν\nu\in\mathbb{R}, α:=CβΓ(2β1)k2β\alpha:=C_{\beta}\Gamma(2\beta-1)k^{-2\beta}. The Lemma 5.4 of [7] ensures that

|Φ¯(ν)(Zz)|CT.\left|\overline{\Phi}(\nu)-\mathbb{P}(Z\leq z)\right|\leq\frac{C}{\sqrt{T}}.

Combining with Lemma 15, we obtain the desired result. ∎

The following Lemma provides the upper bound of D(ν)D(\nu).

Lemma 15.

When TT is large enough, there exists constant Cβ,VC_{\beta,V}^{\prime} such that

supν|D(ν)|Cβ,VTm,\displaystyle\sup_{\nu\in\mathbb{R}}\left|D(\nu)\right|\leq\frac{C_{\beta,V}^{\prime}}{T^{m}},

where m=min{1/3,(34β)/2}m=\min\ \{1/3,(3-4\beta)/2\}.

Proof.

Since the Normal distribution is symmetric, we have

D(ν)\displaystyle D(\nu) =|(kTα2σβ2[1T(0TVt2dt(1T0TVtdt)2)α]ν)Φ¯(ν)|\displaystyle=\left|\mathbb{P}\biggl{(}\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}\biggl{[}\frac{1}{T}\bigl{(}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}\bigr{)}-\alpha\biggr{]}\geq\nu\biggr{)}-\overline{\Phi}(\nu)\right|
=|(kTα2σβ2[1T(0TVt2dt(1T0TVtdt)2)α]ν)Φ(ν)|.\displaystyle=\left|\mathbb{P}\biggl{(}\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}\biggl{[}\frac{1}{T}\bigl{(}\int_{0}^{T}V^{2}_{t}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}\bigr{)}-\alpha\biggr{]}\leq\nu\biggr{)}-\Phi(\nu)\right|.

Consider the following processes:

Xt\displaystyle X_{t} =0tek(ts)dGs,I5=1T(0TXt2dt),\displaystyle=\int_{0}^{t}\mathrm{e}^{-k(t-s)}\mathop{}\!\mathrm{d}G_{s},\;\;\;I_{5}=\frac{1}{T}\biggl{(}\int_{0}^{T}X_{t}^{2}\mathop{}\!\mathrm{d}t\biggr{)}, (27)
FT\displaystyle F_{T} =0T0tek(ts)dGsdt,\displaystyle=\int_{0}^{T}\int_{0}^{t}\mathrm{e}^{-k(t-s)}\mathop{}\!\mathrm{d}G_{s}\mathop{}\!\mathrm{d}t,

where XtX_{t} is an OU process driven by GtG_{t}. According to [11] formula (63), we can obtain

1T0TVt2dt(1T0TVtdt)2α=1T(0TXt2dt)α+KT+I4,\displaystyle\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-\biggl{(}\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t\biggr{)}^{2}-\alpha=\frac{1}{T}\biggl{(}\int_{0}^{T}X_{t}^{2}\mathop{}\!\mathrm{d}t\biggr{)}-\alpha+K_{T}+I_{4},

where

KT\displaystyle K_{T} =μ22k1e2kTTμ2(ekT1)2k2T2,\displaystyle=\frac{\mu^{2}}{2k}\cdot\frac{1-\mathrm{e}^{-2kT}}{T}-\frac{\mu^{2}(\mathrm{e}^{-kT}-1)^{2}}{k^{2}T^{2}},
I4\displaystyle I_{4} =μk[ZTTekTMTT+2(1ekT)FTT2]FT2T2.\displaystyle=\frac{\mu}{k}\biggl{[}\frac{Z_{T}}{T}\mathrm{e}^{-kT}-\frac{M_{T}}{T}+2(1-\mathrm{e}^{-kT})\frac{F_{T}}{T^{2}}\biggr{]}-\frac{F_{T}^{2}}{T^{2}}.

Let a=T1/3a=T^{-1/3}. Lemma 10 ensures that

supν|D(ν)|supν|(kTα2σβ2(I5α+KT)ν)(Zν)|+(|kTα2σβ2I4|>T1/3)+T1/32π.\begin{split}\sup_{\nu\in\mathbb{R}}\left|D(\nu)\right|&\leq\sup_{\nu\in\mathbb{R}}\left|{\mathbb{P}\biggl{(}\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}\bigl{(}I_{5}-\alpha+K_{T}\bigr{)}\leq\nu\biggr{)}}-\mathbb{P}(Z\leq\nu)\right|\\ &\ \ \ +\mathbb{P}\biggl{(}\left|\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}I_{4}\right|>T^{-1/3}\biggr{)}+\frac{T^{-1/3}}{\sqrt{2\pi}}.\end{split} (28)

According to [7] Theorem 1.4, we have

supν|(kTα2σβ2(I5α+KT)ν)(Zν)|Ck,βT34β2,\sup_{\nu\in\mathbb{R}}\left|{\mathbb{P}\biggl{(}\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}\bigl{(}I_{5}-\alpha+K_{T}\bigr{)}\leq\nu\biggr{)}}-\mathbb{P}(Z\leq\nu)\right|\leq\frac{C_{k,\beta}}{T^{\frac{3-4\beta}{2}}}, (29)

where Ck,βC_{k,\beta} independent of TT is a constant. Denote K1=4kα2σβ2K_{1}=4\sqrt{\frac{k}{\alpha^{2}\sigma_{\beta}^{2}}}. We then consider the second term of right side of (28).

(|kTα2σβ2I4|>T1/3)\displaystyle\mathbb{P}\biggl{(}\left|\sqrt{\frac{kT}{\alpha^{2}\sigma_{\beta}^{2}}}I_{4}\right|>T^{-1/3}\biggr{)} (|K1μekTZTk|>T1/6)\displaystyle\leq\mathbb{P}\biggl{(}\left|K_{1}\frac{\mu\cdot\mathrm{e}^{-kT}\cdot Z_{T}}{k}\right|>T^{1/6}\biggr{)}
+(|K1μMTk|>T1/6)\displaystyle\ \ \ +\mathbb{P}\biggl{(}\left|K_{1}\frac{\mu\cdot M_{T}}{k}\right|>T^{1/6}\biggr{)}
+(|K1μ(22ekT)FTk|>T7/6)\displaystyle\ \ \ +\mathbb{P}\biggl{(}\left|K_{1}\frac{\mu\cdot(2-2\mathrm{e}^{-kT})\cdot F_{T}}{k}\right|>T^{7/6}\biggr{)}
+(|K1FT2|>T7/6).\displaystyle\ \ \ +\mathbb{P}\biggl{(}\left|K_{1}\cdot F^{2}_{T}\right|>T^{7/6}\biggr{)}.

Combining the Chebyshev inequality, Lemma 14 and the Proposition 3.10 of [11], we can obtain

(|K1μekTZTk|>T1/6)\displaystyle\mathbb{P}\biggl{(}\left|K_{1}\frac{\mu\cdot\mathrm{e}^{-kT}\cdot Z_{T}}{k}\right|>T^{1/6}\biggr{)} C1𝔼(ZT2)T1/3C1T1/3,\displaystyle\leq\frac{C^{\prime}_{1}\mathbb{E}(Z_{T}^{2})}{T^{1/3}}\leq\frac{C_{1}}{T^{1/3}},
(|K1μMTk|>T1/6)\displaystyle\mathbb{P}\biggl{(}\left|K_{1}\frac{\mu\cdot M_{T}}{k}\right|>T^{1/6}\biggr{)} C2𝔼(MT2)T1/3C2T1/3,\displaystyle\leq\frac{C^{\prime}_{2}\mathbb{E}(M_{T}^{2})}{T^{1/3}}\leq\frac{C_{2}}{T^{1/3}},
(|K1μ(22ekT)FTk|>T13/6)\displaystyle\mathbb{P}\biggl{(}\left|K_{1}\frac{\mu\cdot(2-2\mathrm{e}^{-kT})\cdot F_{T}}{k}\right|>T^{13/6}\biggr{)} C3𝔼(FT2)T13/3C3T13/32β,\displaystyle\leq\frac{C^{\prime}_{3}\mathbb{E}(F_{T}^{2})}{T^{13/3}}\leq\frac{C_{3}}{T^{13/3-2\beta}},
(|K1FT2|>T13/6)\displaystyle\mathbb{P}\Bigl{(}\left|K_{1}\cdot F^{2}_{T}\right|>T^{13/6}\Bigr{)} C4𝔼(FT2)T13/6C4T13/62β.\displaystyle\leq\frac{C^{\prime}_{4}\mathbb{E}(F_{T}^{2})}{T^{13/6}}\leq\frac{C_{4}}{T^{13/6-2\beta}}.

Then we have

(|kTa2σβ2I4|>T1/3)+T1/32πCk,βT1/3,\mathbb{P}\biggl{(}\left|\sqrt{\frac{kT}{a^{2}\sigma_{\beta}^{2}}}I_{4}\right|>T^{-1/3}\biggr{)}+\frac{T^{-1/3}}{\sqrt{2\pi}}\leq\frac{C_{k,\beta}^{\prime}}{T^{1/3}}, (30)

where Ck,βC_{k,\beta}^{\prime} is a constant. Combining formulas (29) and (30), we obtain the desired result. ∎

4. Berry-Esséen upper bounds of least squares estimators

For the convenience of following proof, we introduce some variables:

aT\displaystyle a_{T} =1ekT,bT=1T0Tek(t)𝟙[0,t]()2dtCβΓ(2β1)k2β,\displaystyle=1-\mathrm{e}^{-kT},\;\;\;b_{T}=\frac{1}{T}\int_{0}^{T}\left\|\mathrm{e}^{-k(t-\cdot)}\mathbbm{1}_{[0,t]}(\cdot)\right\|_{\mathfrak{H}}^{2}\mathop{}\!\mathrm{d}t\to C_{\beta}\Gamma(2\beta-1)k^{-2\beta},
cT\displaystyle c_{T} =0Tμ2(1ekt)2dt=μ2(T+2k(ekT1)+12k(1e2kT)),\displaystyle=\int_{0}^{T}\mu^{2}(1-\mathrm{e}^{-kt})^{2}\mathop{}\!\mathrm{d}t=\mu^{2}(T+\frac{2}{k}(\mathrm{e}^{-kT}-1)+\frac{1}{2k}(1-\mathrm{e}^{-2kT})),
dT\displaystyle d_{T} =0T(1ekt)dt=T+1k(ekT1).\displaystyle=\int_{0}^{T}(1-\mathrm{e}^{-kt})\mathop{}\!\mathrm{d}t=T+\frac{1}{k}(\mathrm{e}^{-kT}-1).

The Proposition 3.10 of [11] and Proposition 9 ensure that

eT\displaystyle e_{T} =lT1lT2CT2β,\displaystyle=\left\|l_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}\leq CT^{2\beta},
qT\displaystyle q_{T} =lT1kT2kT11k2CT2β,\displaystyle=\left\|l_{T}\otimes_{1}k_{T}\right\|_{\mathfrak{H}}^{2}\leq\left\|k_{T}\otimes_{1}\frac{1}{k}\right\|_{\mathfrak{H}}^{2}\leq CT^{2\beta},

where CC is a constant independent of TT. Also, we show lTl_{T} and other functions:

fT(t,s)\displaystyle f_{T}(t,s) =ek|ts|𝟙[0,T]2(t,s),\displaystyle=\mathrm{e}^{-k\left|t-s\right|}\mathbbm{1}_{[0,T]^{2}}(t,s),\;\; hT(t,s)=ek(Tt)k(Ts)𝟙[0,T]2(t,s),\displaystyle h_{T}(t,s)=\mathrm{e}^{-k(T-t)-k(T-s)}\mathbbm{1}_{[0,T]^{2}}(t,s),
gT(t,s)\displaystyle g_{T}(t,s) =12kTfThT,\displaystyle=\frac{1}{2kT}f_{T}-h_{T},
kT(s)\displaystyle k_{T}(s) =ek(Ts)𝟙[0,T](s),\displaystyle=\mathrm{e}^{-k(T-s)}\mathbbm{1}_{[0,T]}(s), lT(s)=1k(1ek(Ts))𝟙[0,T](s),\displaystyle l_{T}(s)=\frac{1}{k}(1-\mathrm{e}^{-k(T-s)})\mathbbm{1}_{[0,T]}(s),
mT(s)\displaystyle m_{T}(s) =eks𝟙[0,T](s),\displaystyle=\mathrm{e}^{-ks}\mathbbm{1}_{[0,T]}(s), nT(s)=12k(ek(2Ts)1)𝟙[0,T](s).\displaystyle n_{T}(s)=\frac{1}{2k}(\mathrm{e}^{-k(2T-s)}-1)\mathbbm{1}_{[0,T]}(s).

Furthermore, we denote I1(fT)I_{1}(f_{T}) as I1(fT(t,)𝟙[0,T]())I_{1}(f_{T}(t,\cdot)\mathbbm{1}_{[0,T]}(\cdot)).

We now extent the Corollary 1 of [17].

Theorem 16.

Let FT/HTF_{T}/H_{T} be a zero-mean ratio process that contains at most triple Wiener-Itô stochastic integrals, where FT=I1(f1,T)+I2(f2,T)+I3(f3,T)F_{T}=I_{1}(f_{1,T})+I_{2}(f_{2,T})+I_{3}(f_{3,T}), HT=ET+I1(h1,T)+I2(h2,T)+I3(h3,T)H_{T}=E_{T}+I_{1}(h_{1,T})+I_{2}(h_{2,T})+I_{3}(h_{3,T}) and ETE_{T} is a positive function of TT converging to a constant α\alpha. Suppose that Ψi(T)0\Psi_{i}(T)\to 0, i=1,,4i=1,\cdots,4, as TT\to\infty. Then, there exists constant CC such that when TT is large enough,

supz|(FTHTz)(Zz)|Cmaxp,q=f,hm,n=1,2,3i=1,,(mn)(pm,T~iqn,T(m+n2i),pm,T,qm,Tm,(ET2j=1mj!fj,Tj2)).\begin{split}\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\frac{F_{T}}{H_{T}}\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right|&\leq C\cdot\max_{\begin{subarray}{c}p,q=f,h\\ m,n=1,2,3\\ i=1,\cdots,(m\wedge n)\end{subarray}}\Bigl{(}\left\|p_{m,T}\tilde{\otimes}_{i}q_{n,T}\right\|_{\mathfrak{H}^{\otimes(m+n-2i)}},\\ &\;\;\;\left\langle p_{m,T},q_{m,T}\right\rangle_{\mathfrak{H}^{\otimes m}},(E_{T}^{2}-\sum_{j=1}^{m}j!\left\|f_{j,T}\right\|_{\mathfrak{H}^{\otimes j}}^{2})\Bigr{)}.\end{split} (31)
Proof.

We first consider Ψ4(T)=1(𝔼HT)2𝔼[DHT,DL1(HT𝔼HT)2]\Psi_{4}(T)=\frac{1}{(\mathbb{E}H_{T})^{2}}\sqrt{\mathbb{E}[\left\langle DH_{T},-DL^{-1}(H_{T}-\mathbb{E}H_{T})\right\rangle^{2}_{\mathfrak{H}}]}. It is easy to see that (𝔼HT)2ET2α2a.s.(\mathbb{E}H_{T})^{2}\to E_{T}^{2}\to\alpha^{2}\;\mathrm{a.s.}. Then we deal with 𝔼[DHT,DL1(HT𝔼HT)2]\sqrt{\mathbb{E}[\left\langle DH_{T},-DL^{-1}(H_{T}-\mathbb{E}H_{T})\right\rangle^{2}_{\mathfrak{H}}]}. Denote DHT,DL1(HT𝔼HT)=ψ4\left\langle DH_{T},-DL^{-1}(H_{T}-\mathbb{E}H_{T})\right\rangle_{\mathfrak{H}}=\psi_{4}, we have

ψ4\displaystyle\psi_{4} =h1,T+2I1(h2,T)+3I2(h3,T),h1,T+I1(h2,T)+I2(h3,T)\displaystyle=\left\langle h_{1,T}+2I_{1}(h_{2,T})+3I_{2}(h_{3,T}),h_{1,T}+I_{1}(h_{2,T})+I_{2}(h_{3,T})\right\rangle_{\mathfrak{H}}
=h1,T2+3h1,TI1(h2,T)+4h1,TI2(h3,T)\displaystyle=\left\|h_{1,T}\right\|_{\mathfrak{H}}^{2}+3h_{1,T}I_{1}(h_{2,T})+4h_{1,T}I_{2}(h_{3,T})
+2h2,T22+2I2(h2,T~1h2,T)\displaystyle\;\;\;+2\left\|h_{2,T}\right\|_{\mathfrak{H}^{\otimes 2}}^{2}+2I_{2}(h_{2,T}\tilde{\otimes}_{1}h_{2,T}) (32)
+5I3(h2,T~1h3,T)+10I1(h2,T~2h3,T)\displaystyle\;\;\;+5I_{3}(h_{2,T}\tilde{\otimes}_{1}h_{3,T})+10I_{1}(h_{2,T}\tilde{\otimes}_{2}h_{3,T})
+6h3,T22+3I4(h3,T~1h3,T)+12I2(h3,T~2h3,T).\displaystyle\;\;\;+6\left\|h_{3,T}\right\|_{\mathfrak{H}^{\otimes 2}}^{2}+3I_{4}(h_{3,T}\tilde{\otimes}_{1}h_{3,T})+12I_{2}(h_{3,T}\tilde{\otimes}_{2}h_{3,T}).

Following from the the orthogonality property of multiple integrals, we can obtain

Ψ4(T)\displaystyle\Psi_{4}(T) C1maxm,n=1,2,3i=1,,(mn)(hm,T~ihn,T(m+n2i),hm,Tm2),\displaystyle\leq C_{1}\cdot\max_{\begin{subarray}{c}m,n=1,2,3\\ i=1,\cdots,(m\wedge n)\end{subarray}}\Bigl{(}\left\|h_{m,T}\tilde{\otimes}_{i}h_{n,T}\right\|_{\mathfrak{H}^{\otimes(m+n-2i)}},\left\|h_{m,T}\right\|_{\mathfrak{H}^{\otimes m}}^{2}\Bigr{)}, (33)

where C1C_{1} is a constant independent of TT.

We next consider Ψ2(T)\Psi_{2}(T) and Ψ3(T)\Psi_{3}(T). Denote DFT,DL1(HT𝔼HT)=ψ2\left\langle DF_{T},-DL^{-1}(H_{T}-\mathbb{E}H_{T})\right\rangle_{\mathfrak{H}}=\psi_{2}, we have

ψ2\displaystyle\psi_{2} =f1,T+2I1(f2,T)+3I2(f3,T),h1,T+I1(h2,T)+I2(h3,T)\displaystyle=\left\langle f_{1,T}+2I_{1}(f_{2,T})+3I_{2}(f_{3,T}),h_{1,T}+I_{1}(h_{2,T})+I_{2}(h_{3,T})\right\rangle_{\mathfrak{H}}
=f1,Th1,T+f1,TI1(h2,T)+f1,TI2(h3,T)\displaystyle=f_{1,T}h_{1,T}+f_{1,T}I_{1}(h_{2,T})+f_{1,T}I_{2}(h_{3,T})
+2h1,TI1(f2,T)+3h1,TI2(f3,T)\displaystyle\;\;\;+2h_{1,T}I_{1}(f_{2,T})+3h_{1,T}I_{2}(f_{3,T})
+2f2,T,h2,T2+2I2(f2,T~1h2,T)\displaystyle\;\;\;+2\left\langle f_{2,T},h_{2,T}\right\rangle_{\mathfrak{H}^{\otimes 2}}+2I_{2}(f_{2,T}\tilde{\otimes}_{1}h_{2,T})
+2I3(f2,T~1h3,T)+4I1(f2,T~2h3,T)\displaystyle\;\;\;+2I_{3}(f_{2,T}\tilde{\otimes}_{1}h_{3,T})+4I_{1}(f_{2,T}\tilde{\otimes}_{2}h_{3,T})
+3I3(f3,T~1h2,T)+6I1(f3,T~2h2,T)\displaystyle\;\;\;+3I_{3}(f_{3,T}\tilde{\otimes}_{1}h_{2,T})+6I_{1}(f_{3,T}\tilde{\otimes}_{2}h_{2,T})
+6f3,T,h3,T3+3I4(f3,T~1h3,T)+12I2(f3,T~2h3,T).\displaystyle\;\;\;+6\left\langle f_{3,T},h_{3,T}\right\rangle_{\mathfrak{H}^{\otimes 3}}+3I_{4}(f_{3,T}\tilde{\otimes}_{1}h_{3,T})+12I_{2}(f_{3,T}\tilde{\otimes}_{2}h_{3,T}).

Simalarly, there exists a constant C2C_{2} such that

Ψ2(T)\displaystyle\Psi_{2}(T) C2maxp,q=f,hm,n=1,2,3i=1,,(mn)(pm,T~iqn,T(m+n2i),fm,T,hm,Tm).\displaystyle\leq C_{2}\cdot\max_{\begin{subarray}{c}p,q=f,h\\ m,n=1,2,3\\ i=1,\cdots,(m\wedge n)\end{subarray}}\Bigl{(}\left\|p_{m,T}\tilde{\otimes}_{i}q_{n,T}\right\|_{\mathfrak{H}^{\otimes(m+n-2i)}},\left\langle f_{m,T},h_{m,T}\right\rangle_{\mathfrak{H}^{\otimes m}}\Bigr{)}. (34)

Following from the above result, we can obtain the bound of Ψ3(T)\Psi_{3}(T).

We then deal with Ψ1(T)=1(𝔼HT)2𝔼[((𝔼HT)2DFT,DL1FT)2]\Psi_{1}(T)=\frac{1}{(\mathbb{E}{H_{T}})^{2}}\sqrt{\mathbb{E}[((\mathbb{E}H_{T})^{2}-\left\langle DF_{T},-DL^{-1}F_{T}\right\rangle_{\mathfrak{H}})^{2}]}. According to the orthogonality and (32), we have

Ψ1(T)\displaystyle\Psi_{1}(T) C3maxm,n=1,2,3i=1,,(mn)(fm,T~ifn,T(m+n2i),(ET2i=1mj!fj,Tj2)).\displaystyle\leq C_{3}\cdot\max_{\begin{subarray}{c}m,n=1,2,3\\ i=1,\cdots,(m\wedge n)\end{subarray}}\Bigl{(}\left\|f_{m,T}\tilde{\otimes}_{i}f_{n,T}\right\|_{\mathfrak{H}^{\otimes(m+n-2i)}},(E_{T}^{2}-\sum_{i=1}^{m}j!\left\|f_{j,T}\right\|_{\mathfrak{H}^{\otimes j}}^{2})\Bigr{)}. (35)

Combining formulas (33), (34) and (35), we obtain the desired result. ∎

We now prove the convergence speed of k^LS\hat{k}_{LS}. First, we review the CLT of k^LS\hat{k}_{LS}.

Theorem 17 ([11] Propositions 4.20).

When β(1/2,3/4)\beta\in(1/2,3/4) and 1 holds, k^LS\hat{k}_{LS} satisfies the following central limit theorem:

T(k^LSk)\displaystyle\sqrt{T}(\hat{k}_{LS}-k) =VTTμ^+kTμ^(μ^μ)1T0TVtdGt1T0TVt2dt(1T0TVtdt)2law𝒩(0,kσβ2).\displaystyle=\frac{\frac{V_{T}}{\sqrt{T}}\hat{\mu}+k\sqrt{T}\hat{\mu}(\hat{\mu}-\mu)-\frac{1}{\sqrt{T}}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}G_{t}}{\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}}\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,k\sigma_{\beta}^{2}). (36)

We then transfrom the above T(k^LSk)\sqrt{T}(\hat{k}_{LS}-k) as multiple Wiener-Itô integrals.

Proposition 18.

Let Xt=0tek(ts)dGsX_{t}=\int_{0}^{t}\mathrm{e}^{-k(t-s)}\mathop{}\!\mathrm{d}G_{s} be a OU process driven by GtG_{t}. T(k^LSk)\sqrt{T}(\hat{k}_{LS}-k) can be rewritten as:

T(k^LSk)=I0+I1(f1,T)+I2(f2,T)J0+I1(h1,T)+J2(h2,T).\sqrt{T}(\hat{k}_{LS}-k)=\frac{I_{0}+I_{1}(f_{1,T})+I_{2}(f_{2,T})}{J_{0}+I_{1}(h_{1,T})+J_{2}(h_{2,T})}. (37)

where

I0\displaystyle I_{0} =1T3/2(μ2aTdT+qT+μ2kaT2+keT)+μ2T1/2(ekT1),\displaystyle=\frac{1}{T^{3/2}}\Bigl{(}\mu^{2}a_{T}\cdot d_{T}+q_{T}+\frac{\mu^{2}}{k}a_{T}^{2}+ke_{T}\Bigr{)}+\frac{\mu^{2}}{T^{1/2}}(\mathrm{e}^{-kT}-1), (38a)
f1,T\displaystyle f_{1,T} =μT3/2(dTkTaTlT)+μT(mTkT),\displaystyle=\frac{\mu}{T^{3/2}}\bigl{(}d_{T}k_{T}-a_{T}l_{T}\bigr{)}+\frac{\mu}{\sqrt{T}}\bigl{(}m_{T}-k_{T}\bigr{)}, (38b)
f2,T\displaystyle f_{2,T} =1T3/2lTkT+kT3/2lTlT12TfT,\displaystyle=\frac{1}{T^{3/2}}l_{T}\otimes k_{T}+\frac{k}{T^{3/2}}l_{T}\otimes l_{T}-\frac{1}{2\sqrt{T}}f_{T}, (38c)
J0\displaystyle J_{0} =cTT+bT1T2(μ2dT2+eT),\displaystyle=\frac{c_{T}}{T}+b_{T}-\frac{1}{T^{2}}(\mu^{2}d_{T}^{2}+e_{T}), (38d)
h1,T\displaystyle h_{1,T} =2μT(lT+nT)2T2dTlT,\displaystyle=\frac{2\mu}{T}\bigl{(}l_{T}+n_{T}\bigr{)}-\frac{2}{T^{2}}d_{T}l_{T}, (38e)
h2,T\displaystyle h_{2,T} =gT1T2lTlT.\displaystyle=g_{T}-\frac{1}{T^{2}}l_{T}\otimes l_{T}. (38f)
Proof.

We first deal with the numerator of (36):

VTTμ^+kTμ^(μ^μ)1T0TVtdGt.\frac{V_{T}}{\sqrt{T}}\hat{\mu}+k\sqrt{T}\hat{\mu}(\hat{\mu}-\mu)-\frac{1}{\sqrt{T}}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}G_{t}. (39)

According to the definition of above functions, VT=μaT+I1(kT)V_{T}=\mu a_{T}+I_{1}(k_{T}). Combining with Proposition 9, we can obtain

VTTμ^\displaystyle\frac{V_{T}}{\sqrt{T}}\hat{\mu} =μaT+I1(kT)T1/2μdT+I1(lT)T\displaystyle=\frac{\mu a_{T}+I_{1}(k_{T})}{T^{1/2}}\frac{\mu d_{T}+I_{1}(l_{T})}{T}
=1T3/2(μ2aTdT+μaTI1(lT)+μdTI1(kT)+I1(lT)I1(kT))\displaystyle=\frac{1}{T^{3/2}}\bigl{(}\mu^{2}a_{T}\cdot d_{T}+\mu a_{T}I_{1}(l_{T})+\mu d_{T}I_{1}(k_{T})+I_{1}(l_{T})I_{1}(k_{T})\bigr{)}
=1T3/2(μ2aTdT+μaTI1(lT)+μdTI1(kT)\displaystyle=\frac{1}{T^{3/2}}\Bigl{(}\mu^{2}a_{T}\cdot d_{T}+\mu a_{T}I_{1}(l_{T})+\mu d_{T}I_{1}(k_{T})
+I2(lTkT)+lT1kT2).\displaystyle\ \ \ +I_{2}(l_{T}\otimes k_{T})+\left\|l_{T}\otimes_{1}k_{T}\right\|_{\mathfrak{H}}^{2}\Bigr{)}.

Then the second term of (39). Let Item2:=kTμ^(μ^μ)Item_{2}:=k\sqrt{T}\hat{\mu}(\hat{\mu}-\mu), we have

Item2\displaystyle Item_{2} =kT(μ^μ+μ)(μ^μ)\displaystyle=k\sqrt{T}(\hat{\mu}-\mu+\mu)(\hat{\mu}-\mu)
=kT(μ^μ)2+kTμ(μ^μ)\displaystyle=k\sqrt{T}(\hat{\mu}-\mu)^{2}+k\sqrt{T}\mu(\hat{\mu}-\mu)
=kT3/2(μ2k2(ekT1)2+2μk(ekT1)I1(lT)+I2(lTlT)+eT)\displaystyle=\frac{k}{T^{3/2}}\Bigl{(}\frac{\mu^{2}}{k^{2}}(\mathrm{e}^{-kT}-1)^{2}+\frac{2\mu}{k}(\mathrm{e}^{-kT}-1)I_{1}(l_{T})+I_{2}(l_{T}\otimes l_{T})+e_{T}\Bigr{)}
+kμT1/2(μk(ekT1)+I1(lT))\displaystyle\ \ \ +\frac{k\mu}{T^{1/2}}\Bigl{(}\frac{\mu}{k}(\mathrm{e}^{-kT}-1)+I_{1}(l_{T})\Bigr{)}
=kT3/2(μ2k2(ekT1)2+2μk(ekT1)I1(lT)+I2(lTlT)+eT)\displaystyle=\frac{k}{T^{3/2}}\Bigl{(}\frac{\mu^{2}}{k^{2}}(\mathrm{e}^{-kT}-1)^{2}+\frac{2\mu}{k}(\mathrm{e}^{-kT}-1)I_{1}(l_{T})+I_{2}(l_{T}\otimes l_{T})+e_{T}\Bigr{)}
+μ2T1/2(ekT1)+μT1/2(GTI1(kT)).\displaystyle\ \ \ +\frac{\mu^{2}}{T^{1/2}}(\mathrm{e}^{-kT}-1)+\frac{\mu}{T^{1/2}}\bigl{(}G_{T}-I_{1}(k_{T})\bigr{)}.

Besides, we can obtain

1T0TVtdGt\displaystyle-\frac{1}{\sqrt{T}}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}G_{t} =1T(0Tμ(1ekt)dGt+12I2(fT))\displaystyle=-\frac{1}{\sqrt{T}}\Bigl{(}\int_{0}^{T}\mu(1-\mathrm{e}^{-kt})\mathop{}\!\mathrm{d}G_{t}+\frac{1}{2}I_{2}(f_{T})\Bigr{)}
=1T(μ(GTI1(mT))+12I2(fT)).\displaystyle=-\frac{1}{\sqrt{T}}\Bigl{(}\mu(G_{T}-I_{1}(m_{T}))+\frac{1}{2}I_{2}(f_{T})\Bigr{)}.

Next, we consider the denominator. For 1T0TVt2dt\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t, we have

1T0TVt2dt\displaystyle\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t =1T0T(μ(1ekt)+Xt)2dt\displaystyle=\frac{1}{T}\int_{0}^{T}(\mu(1-\mathrm{e}^{-kt})+X_{t})^{2}\mathop{}\!\mathrm{d}t
=1T(0Tμ2(1ekt)2dt+20tμ(1ekt)Xtdt+0TXt2dt)\displaystyle=\frac{1}{T}\Bigl{(}\int_{0}^{T}\mu^{2}(1-\mathrm{e}^{-kt})^{2}\mathop{}\!\mathrm{d}t+2\int_{0}^{t}\mu(1-\mathrm{e}^{-kt})X_{t}\mathop{}\!\mathrm{d}t+\int_{0}^{T}X_{t}^{2}\mathop{}\!\mathrm{d}t\Bigr{)}
=1T(cT+2μ0T0t(1ekt)ek(ts)dGsdt)+I2(gT)+bT\displaystyle=\frac{1}{T}\Bigl{(}c_{T}+2\mu\int_{0}^{T}\int_{0}^{t}(1-\mathrm{e}^{-kt})\mathrm{e}^{-k(t-s)}\mathop{}\!\mathrm{d}G_{s}\mathop{}\!\mathrm{d}t\Bigr{)}+I_{2}(g_{T})+b_{T}
=cTT+2μT(I1(lT)+I1(nT))+I2(gT)+bT.\displaystyle=\frac{c_{T}}{T}+\frac{2\mu}{T}\bigl{(}I_{1}(l_{T})+I_{1}(n_{T})\bigr{)}+I_{2}(g_{T})+b_{T}.

Since μ^2=(1T0TVtdt)2\hat{\mu}^{2}=(\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t)^{2}, we can obtain

(1T0TVtdt)2\displaystyle\Bigl{(}\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t\Bigr{)}^{2} =1T2(μdT+I1(lT))2\displaystyle=\frac{1}{T^{2}}\Bigl{(}\mu d_{T}+I_{1}(l_{T})\Bigr{)}^{2}
=1T2(μ2dT2+2μdTI1(lT)+I12(lT))\displaystyle=\frac{1}{T^{2}}\Bigl{(}\mu^{2}d_{T}^{2}+2\mu d_{T}I_{1}(l_{T})+I_{1}^{2}(l_{T})\Bigr{)}
=1T2(μ2dT2+2μdTI1(lT)+I2(lTlT)+eT).\displaystyle=\frac{1}{T^{2}}\Bigl{(}\mu^{2}d_{T}^{2}+2\mu d_{T}I_{1}(l_{T})+I_{2}(l_{T}\otimes l_{T})+e_{T}\Bigr{)}.

Combining the above formulas, we obtain the desired result. ∎

For simplicity, let FT:=(I1(f1,T)+I2(f2,T))/(kσβ),HT:=J0+I1(h1,T)+I2(h2,T)F_{T}:=(I_{1}(f_{1,T})+I_{2}(f_{2,T}))/(\sqrt{k}\sigma_{\beta}),\;H_{T}:=J_{0}+I_{1}(h_{1,T})+I_{2}(h_{2,T}). We show the convergence speed of zero-mean part.

Lemma 19.

Let Z𝒩(0,1)Z\sim\mathcal{N}(0,1) be a standard Normal variable. Assume β(1/2,3/4)\beta\in(1/2,3/4) and 1 holds. When TT is large enough, there exists constant Cβ,VC_{\beta,V}^{\prime} such that

supz|(FTHTz)(Zz)|Cβ,VTγ,\sup_{z\in\mathbb{R}}\left|\mathbb{P}\left(\frac{F_{T}}{H_{T}}\leq z\right)-\mathbb{P}(Z\leq z)\right|\leq\frac{C_{\beta,V}^{\prime}}{T^{\gamma}}, (40)

where γ=min{1/2,34β}\gamma=\min{\{1/2,3-4\beta\}}.

Proof.

According to [11] formulas (9) and (47),

𝔼HT=𝔼J0α=CβΓ(2β1)k2βa.s..\displaystyle\mathbb{E}H_{T}=\mathbb{E}J_{0}\to\alpha=C_{\beta}\Gamma(2\beta-1)k^{-2\beta}\;\;\;\mathrm{a.s.}.

Combining with [12] Lemma 5.2.4, we can obtain

max{lT2,nT2}C(1)T2β,\displaystyle\max\{\left\|l_{T}\right\|_{\mathfrak{H}}^{2},\left\|n_{T}\right\|_{\mathfrak{H}}^{2}\}\leq C^{(1)}T^{2\beta}, (41)

where C(1)C^{(1)} is a constant, and h1,T2C(2)/T22β\left\|h_{1,T}\right\|_{\mathfrak{H}}^{2}\leq C^{(2)}/T^{2-2\beta}.

Following from [7] Theorem 1.4 and (41), we have

gT2C(3)T,gT1gT2C(3)T,2T2lTlT22C(3)T44β,\displaystyle\left\|g_{T}\right\|_{\mathfrak{H}^{\otimes 2}}\leq\frac{C^{(3)}}{\sqrt{T}},\;\;\;\left\|g_{T}\otimes_{1}g_{T}\right\|_{\mathfrak{H}^{\otimes 2}}\leq\frac{C^{(3)}}{T},\;\;\;\left\|\frac{2}{T^{2}}l_{T}\otimes l_{T}\right\|_{\mathfrak{H}^{\otimes 2}}^{2}\leq\frac{C^{(3)}}{T^{4-4\beta}},

where C(3)C^{(3)} is a constant independent of TT. Minkowski inequality ensures that h2,T22C(4)/T\left\|h_{2,T}\right\|_{\mathfrak{H}^{\otimes 2}}^{2}\leq C^{(4)}/T. Simalarly, the Proposition 3.10 of [11] induces that

J02j=1mj!fj,Tj2kσβ2C(5)/T34β,\displaystyle J_{0}^{2}-\sum_{j=1}^{m}\frac{j!\left\|f_{j,T}\right\|_{\mathfrak{H}^{\otimes j}}^{2}}{k\sigma_{\beta}^{2}}\leq C^{(5)}/T^{3-4\beta}, (42)

and f1,T2C(5)/T\left\|f_{1,T}\right\|_{\mathfrak{H}}^{2}\leq C^{(5)}/T. The formula (5.13) of[7] ensures that

f2,T~1h2,T2C(5)/T1/2.\displaystyle\left\|f_{2,T}\tilde{\otimes}_{1}h_{2,T}\right\|_{\mathfrak{H}^{\otimes 2}}\leq C^{(5)}/T^{1/2}.

Combining the above results, Theorem 16 and Cauchy-Schwarz inequality, we obtain the Lemma. ∎

The following Lemma shows the bound of non-zero mean part.

Lemma 20.

Assume β(1/2,3/4)\beta\in(1/2,3/4) and 1 holds. When TT is large enough, there exists constant C1C_{1} such that

(|L1|>C1T(3/4β))=0,\mathbb{P}\Bigl{(}\left|L_{1}\right|>\frac{C_{1}}{T^{(3/4-\beta)}}\Bigr{)}=0,

where L1=I0/(kσβHT)L_{1}=I_{0}/(\sqrt{k}\sigma_{\beta}H_{T}).

Proof.

The Proposition 3.14 and Corollary 3.15 of [11] ensure that when TT is large enough,

1T0TVt2dt(1T0TVtdt)2CβΓ(2β1)k2β=αa.s..\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-\Bigl{(}\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t\Bigr{)}^{2}\to C_{\beta}\Gamma(2\beta-1)k^{-2\beta}=\alpha\;\;\;\mathrm{a.s.}. (43)

Then we can obtain

|L1|=|1kσβI0HT|C|I0|,\left|L_{1}\right|=\left|\frac{\frac{1}{\sqrt{k}\sigma_{\beta}}I_{0}}{H_{T}}\right|\leq C^{\prime}\left|I_{0}\right|,

where CC^{\prime} is a constant. Furthermore, According to (38a), there exists C′′C^{\prime\prime} such that

|I0|1T3/2(μ2aTdT+qT+μ2kaT2+keT)+μ2T1/2aTC′′T3/22β.\left|I_{0}\right|\leq\frac{1}{T^{3/2}}\Bigl{(}\mu^{2}a_{T}\cdot d_{T}+q_{T}+\frac{\mu^{2}}{k}a_{T}^{2}+ke_{T}\Bigr{)}+\frac{\mu^{2}}{T^{1/2}}a_{T}\leq\frac{C^{\prime\prime}}{T^{3/2-2\beta}}.

Combining the above two formulas, we have

|L1|C(6)T3/22βa.s.,\left|L_{1}\right|\leq\frac{C^{(6)}}{T^{3/2-2\beta}}\;\;\;\mathrm{a.s.},

where C(6)=2CC′′C^{(6)}=2C^{\prime}\cdot C^{\prime\prime}. Then we obtain the desired result. ∎

We now prove the formula (9).

Proof of formula (9).

According to Lemma 10,

supz|(Tkσβ2(k^LSk)z)(Zz)|\displaystyle\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\sqrt{\frac{T}{k\sigma_{\beta}^{2}}}(\hat{k}_{LS}-k)\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right| supz|(FTHTz)(Zz)|\displaystyle\leq\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\frac{F_{T}}{H_{T}}\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right|
+(|L1|>C1T3/4β)+12πC1T3/4β.\displaystyle\ \ \ +\mathbb{P}\Bigl{(}\left|L_{1}\right|>\frac{C_{1}}{T^{3/4-\beta}}\Bigr{)}+\frac{1}{\sqrt{2\pi}}\frac{C_{1}}{T^{3/4-\beta}}.

Combining Lemmas 19 and 20, we obtain the desired result. ∎

Pei et al.[11] show the CLT of least squares estimator μ^LS\hat{\mu}_{LS} of mean coefficient μ\mu.

Theorem 21 ([11] Propositions 4.21).

Assume β(1/2,1)\beta\in(1/2,1) and GtG_{t} is a self-similar Gaussian process satisfing 1 and 𝔼[G12]=1\mathbb{E}[G_{1}^{2}]=1. T1β(μ^LSμ)T^{1-\beta}(\hat{\mu}_{LS}-\mu) is asymptotically normal as TT\to\infty:

T1β(μ^LSμ)=VTTβ[1T0TVt2dtμμ^]1T0TVtdVtT1β[μ^μ]VTTμ^1T0TVtdVtlaw𝒩(0,1/k2).T^{1-\beta}(\hat{\mu}_{LS}-\mu)=\frac{\frac{V_{T}}{T^{\beta}}\big{[}\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-\mu\hat{\mu}\big{]}-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}\cdot T^{1-\beta}\big{[}\hat{\mu}-\mu\big{]}}{\frac{V_{T}}{T}\cdot\hat{\mu}-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}}\stackrel{{\scriptstyle law}}{{\longrightarrow}}\mathcal{N}(0,1/k^{2}). (44)

We also transform T1β(μ^LSμ)T^{1-\beta}(\hat{\mu}_{LS}-\mu) as the following multiple integrals.

Proposition 22.

T1β(μ^LSμ)T^{1-\beta}(\hat{\mu}_{LS}-\mu) can be represented by as :

T1β(μ^LSμ)=I0+I1(f1,T)+I2(f2,T)+I3(f3,T)J0+I1(h1,T)+I2(h1,T),T^{1-\beta}(\hat{\mu}_{LS}-\mu)=\frac{I_{0}^{*}+I_{1}(f_{1,T}^{*})+I_{2}(f_{2,T}^{*})+I_{3}(f_{3,T}^{*})}{J_{0}^{*}+I_{1}(h_{1,T}^{*})+I_{2}(h_{1,T}^{*})},

where

I0=1T1+β(2μ(kT1lT2+kT1nT2)+2kμnT1lT2+μmT1lT2),\displaystyle I_{0}^{*}=\frac{1}{T^{1+\beta}}\Bigl{(}2\mu\bigl{(}\left\|k_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}+\left\|k_{T}\otimes_{1}n_{T}\right\|_{\mathfrak{H}}^{2}\bigr{)}+2k\mu\left\|n_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}+\mu\left\|m_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}\Bigr{)}, (45a)
f1,T=1T1+β((cTμ2dTμ2aTk𝟙[0,T])+2μ2aTlT+fT1lT)\displaystyle f_{1,T}^{*}=\frac{1}{T^{1+\beta}}\Bigl{(}(c_{T}-\mu^{2}d_{T}-\frac{\mu^{2}a_{T}}{k}\mathbbm{1}_{[0,T]})+2\mu^{2}a_{T}l_{T}+f_{T}\otimes_{1}l_{T}\Bigr{)}
1T1+βμ2aTkmT+1Tβ(bT𝟙[0,T]+2gT1𝟙[0,T]),\displaystyle\;\;\;\;\;\;-\frac{1}{T^{1+\beta}}\frac{\mu^{2}a_{T}}{k}m_{T}+\frac{1}{T^{\beta}}\Bigl{(}b_{T}\mathbbm{1}_{[0,T]}+2g_{T}\otimes_{1}\mathbbm{1}_{[0,T]}\Bigr{)}, (45b)
f2,T=1T1+β(2μ(kTlT+kTnT)+2kμ(nTlT)+μmTlTμ2kaTfT),\displaystyle f_{2,T}^{*}=\frac{1}{T^{1+\beta}}\Bigl{(}2\mu\bigl{(}k_{T}\otimes l_{T}+k_{T}\otimes n_{T}\bigr{)}+2k\mu(n_{T}\otimes l_{T})+\mu m_{T}\otimes l_{T}-\frac{\mu}{2k}a_{T}f_{T}\Bigr{)}, (45c)
f3,T=1Tβ(gTkT+kgTlT)+12T1+βfTlT,\displaystyle f_{3,T}^{*}=\frac{1}{T^{\beta}}\bigl{(}g_{T}\otimes k_{T}+kg_{T}\otimes l_{T}\bigr{)}+\frac{1}{2T^{1+\beta}}f_{T}\otimes l_{T}, (45d)
J0=1T2(μ2aTdT+lT1kT2)+1T(kcTkμ2dT)+kbT,\displaystyle J_{0}^{*}=\frac{1}{T^{2}}\Bigl{(}\mu^{2}a_{T}\cdot d_{T}+\left\|l_{T}\otimes_{1}k_{T}\right\|_{\mathfrak{H}}^{2}\Bigr{)}+\frac{1}{T}\Bigl{(}kc_{T}-k\mu^{2}d_{T}\bigr{)}+kb_{T}, (45e)
h1,T=1T2(μaTlT+μdTkT)+1T(μkT+2kμnT+μmT),\displaystyle h_{1,T}^{*}=\frac{1}{T^{2}}\Bigl{(}\mu a_{T}l_{T}+\mu d_{T}k_{T}\Bigr{)}+\frac{1}{T}\Bigl{(}\mu k_{T}+2k\mu n_{T}+\mu m_{T}\Bigr{)}, (45f)
h2,T=1T2lTkT12TfT+kgT.\displaystyle h_{2,T}^{*}=\frac{1}{T^{2}}l_{T}\otimes k_{T}-\frac{1}{2T}f_{T}+kg_{T}. (45g)
Proof.

We first consider the denominator.

VTTμ^1T0TVtdVt.\displaystyle\frac{V_{T}}{T}\cdot\hat{\mu}-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}.

According to Proposition 18,

VTTμ^\displaystyle\frac{V_{T}}{T}\hat{\mu} =μaT+I1(kT)TμdT+I1(lT)T\displaystyle=\frac{\mu a_{T}+I_{1}(k_{T})}{T}\frac{\mu d_{T}+I_{1}(l_{T})}{T}
=1T2(μ2aTdT+μaTI1(lT)+μdTI1(kT)+I1(lT)I1(kT))\displaystyle=\frac{1}{T^{2}}\bigl{(}\mu^{2}a_{T}\cdot d_{T}+\mu a_{T}I_{1}(l_{T})+\mu d_{T}I_{1}(k_{T})+I_{1}(l_{T})I_{1}(k_{T})\bigr{)}
=1T2(μ2aTdT+μaTI1(lT)+μdTI1(kT)+I2(lTkT)+lT1kT2).\displaystyle=\frac{1}{T^{2}}\Bigl{(}\mu^{2}a_{T}\cdot d_{T}+\mu a_{T}I_{1}(l_{T})+\mu d_{T}I_{1}(k_{T})+I_{2}(l_{T}\otimes k_{T})+\left\|l_{T}\otimes_{1}k_{T}\right\|_{\mathfrak{H}}^{2}\Bigr{)}.

Then, we deal with 1T0TVtdVt-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}:

1T0TVtdVt\displaystyle-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t} =1T(0TkμVtdt0TkVt2dt+0TVtdGt)\displaystyle=-\frac{1}{T}\Bigl{(}\int_{0}^{T}k\mu V_{t}\mathop{}\!\mathrm{d}t-\int_{0}^{T}kV_{t}^{2}\mathop{}\!\mathrm{d}t+\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}G_{t}\Bigr{)}
=kμT0TVtdt+kT0TVt2dt1T0TVtdGt\displaystyle=-\frac{k\mu}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}t+\frac{k}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}G_{t}
=kμ2dT+kμI1(lT)T+(kcTT+2kμT(I1(lT)+I1(nT)))\displaystyle=-\frac{k\mu^{2}d_{T}+k\mu I_{1}(l_{T})}{T}+\Bigl{(}\frac{kc_{T}}{T}+\frac{2k\mu}{T}\bigl{(}I_{1}(l_{T})+I_{1}(n_{T})\bigr{)}\Bigr{)}
1T(μ(GTI1(mT))+12I2(fT))+kI2(gT)+kbT\displaystyle\ \ \ -\frac{1}{T}\Bigl{(}\mu\bigl{(}G_{T}-I_{1}(m_{T})\bigr{)}+\frac{1}{2}I_{2}(f_{T})\Bigr{)}+kI_{2}(g_{T})+kb_{T}
=1T(kcTkμ2dT+2kμI1(nT)+μI1(kT)+μI1(mT)+12I2(fT))\displaystyle=\frac{1}{T}\Bigl{(}kc_{T}-k\mu^{2}d_{T}+2k\mu I_{1}(n_{T})+\mu I_{1}(k_{T})+\mu I_{1}(m_{T})+\frac{1}{2}I_{2}(f_{T})\Bigr{)}
+kI2(gT)+kbT.\displaystyle\ \ \ +kI_{2}(g_{T})+kb_{T}.

We next consider the numerator:

VTTβ[1T0TVt2dtμμ^]1T0TVtdVtT1β[μ^μ].\displaystyle\frac{V_{T}}{T^{\beta}}\Big{[}\frac{1}{T}\int_{0}^{T}V_{t}^{2}\mathop{}\!\mathrm{d}t-\mu\hat{\mu}\Big{]}-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}\cdot T^{1-\beta}\big{[}\hat{\mu}-\mu\big{]}.

Since the Proposition 18, we have

VTTβ1T0TVt2𝑑t\displaystyle\frac{V_{T}}{T^{\beta}}\frac{1}{T}\int_{0}^{T}V_{t}^{2}dt =1Tβ(μaT+I1(kT))(cTT+2μT(I1(lT)+I1(nT))+I2(gT)+bT)\displaystyle=\frac{1}{T^{\beta}}\bigl{(}\mu a_{T}+I_{1}(k_{T})\bigr{)}\Bigl{(}\frac{c_{T}}{T}+\frac{2\mu}{T}\bigl{(}I_{1}(l_{T})+I_{1}(n_{T})\bigr{)}+I_{2}(g_{T})+b_{T}\Bigr{)}
=1T1+β(μaTcT+2μ2aT(I1(lT)+I1(nT))+cTI1(kT)\displaystyle=\frac{1}{T^{1+\beta}}\Bigl{(}\mu a_{T}\cdot c_{T}+2\mu^{2}a_{T}\bigl{(}I_{1}(l_{T})+I_{1}(n_{T})\bigr{)}+c_{T}I_{1}(k_{T})
+2μ(I2(kTlT)+I2(kTnT))\displaystyle\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+2\mu\bigl{(}I_{2}(k_{T}\otimes l_{T})+I_{2}(k_{T}\otimes n_{T})\bigr{)}
+2μ(kT1lT2+kT1nT2))\displaystyle\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+2\mu\bigl{(}\left\|k_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}+\left\|k_{T}\otimes_{1}n_{T}\right\|_{\mathfrak{H}}^{2}\bigr{)}\Bigr{)}
+1Tβ(μaTI2(gT)+μaTbT+I3(kTgT)\displaystyle\;\;\;+\frac{1}{T^{\beta}}\Bigl{(}\mu a_{T}I_{2}(g_{T})+\mu a_{T}b_{T}+I_{3}(k_{T}\otimes g_{T})
+2I1(kT1gT)+bTI1(kT)).\displaystyle\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+2I_{1}(k_{T}\otimes_{1}g_{T})+b_{T}I_{1}(k_{T})\Bigr{)}.

Similarly, we can obtain

VTTβμμ^\displaystyle-\frac{V_{T}}{T^{\beta}}\cdot\mu\hat{\mu} =μμaT+I1(kT)TβμdT+I1(lT)T\displaystyle=-\mu\frac{\mu a_{T}+I_{1}(k_{T})}{T^{\beta}}\frac{\mu d_{T}+I_{1}(l_{T})}{T}
=μT1+β(μ2aTdT+μaTI1(lT)+μdTI1(kT)+I1(lT)I1(kT))\displaystyle=\frac{-\mu}{T^{1+\beta}}(\mu^{2}a_{T}\cdot d_{T}+\mu a_{T}I_{1}(l_{T})+\mu d_{T}I_{1}(k_{T})+I_{1}(l_{T})I_{1}(k_{T}))
=μT1+β(μ2aTdT+μaTI1(lT)+μdTI1(kT)\displaystyle=\frac{-\mu}{T^{1+\beta}}\Bigl{(}\mu^{2}a_{T}\cdot d_{T}+\mu a_{T}I_{1}(l_{T})+\mu d_{T}I_{1}(k_{T})
+I2(lTkT)+lT1kT2).\displaystyle\ \ \ +I_{2}(l_{T}\otimes k_{T})+\left\|l_{T}\otimes_{1}k_{T}\right\|_{\mathfrak{H}}^{2}\Bigr{)}.

Let Item3=1T0TVtdVtT1β(μ^μ)Item_{3}=-\frac{1}{T}\int_{0}^{T}V_{t}\mathop{}\!\mathrm{d}V_{t}\cdot T^{1-\beta}(\hat{\mu}-\mu), we have

Item3\displaystyle Item_{3} =(I1(lT)TβμaTkTβ)1T0TVtdVt\displaystyle=\Bigl{(}\frac{I_{1}(l_{T})}{T^{\beta}}-\frac{\mu a_{T}}{kT^{\beta}}\Bigr{)}\frac{1}{T}\int_{0}^{T}-V_{t}\mathop{}\!\mathrm{d}V_{t}
=1T1+β(kcTI1(lT)kμ2dTI1(lT)+2kμI2(nTlT)+2kμnT1lT2\displaystyle=\frac{1}{T^{1+\beta}}\biggl{(}kc_{T}I_{1}(l_{T})-k\mu^{2}d_{T}I_{1}(l_{T})+2k\mu I_{2}(n_{T}\otimes l_{T})+2k\mu\left\|n_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}
+μI2(kTlT)+μkT1lT2+μI2(mTlT)+μmT1lT2)\displaystyle\ \ \ +\mu I_{2}(k_{T}\otimes l_{T})+\mu\left\|k_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}+\mu I_{2}(m_{T}\otimes l_{T})+\mu\left\|m_{T}\otimes_{1}l_{T}\right\|_{\mathfrak{H}}^{2}\biggr{)}
+kTβ(bTI1(lT)+I3(gTlT)+2I1(gT1lT))\displaystyle+\frac{k}{T^{\beta}}\Bigl{(}b_{T}I_{1}(l_{T})+I_{3}(g_{T}\otimes l_{T})+2I_{1}(g_{T}\otimes_{1}l_{T})\Bigr{)}
+12T1+β(I3(fTlT)+2I1(fT1lT))\displaystyle+\frac{1}{2T^{1+\beta}}\Bigl{(}I_{3}(f_{T}\otimes l_{T})+2I_{1}(f_{T}\otimes_{1}l_{T})\Bigr{)}
+1T1+β(μ3aTdTμaTcT2μ2aTI1(nT))\displaystyle+\frac{1}{T^{1+\beta}}\Bigl{(}\mu^{3}a_{T}\cdot d_{T}-\mu a_{T}\cdot c_{T}-2\mu^{2}a_{T}I_{1}(n_{T})\Bigr{)}
1kT1+β(μ2aTI1(kT)+μ2aTI1(mT)+12μaTI2(fT))\displaystyle-\frac{1}{kT^{1+\beta}}\Bigl{(}\mu^{2}a_{T}I_{1}(k_{T})+\mu^{2}a_{T}I_{1}(m_{T})+\frac{1}{2}\mu a_{T}I_{2}(f_{T})\Bigr{)}
1Tβ(μaTbT+μaTI2(gT)).\displaystyle-\frac{1}{T^{\beta}}\bigl{(}\mu a_{T}\cdot b_{T}+\mu a_{T}I_{2}(g_{T})\bigr{)}.

Combining the above formulas, we obtain the Proposition. ∎

Denote FT:=I0+I1(f1,T)+I2(f2,T)+I3(f3,T)F_{T}^{*}:=I_{0}^{*}+I_{1}(f_{1,T}^{*})+I_{2}(f_{2,T}^{*})+I_{3}(f_{3,T}^{*}) and HT=1k(J0+I1(h1,T)+I2(h1,T))H_{T}^{*}=\frac{1}{k}(J_{0}^{*}+I_{1}(h_{1,T}^{*})+I_{2}(h_{1,T}^{*})). We now prove the upper bounds of zero-mean part.

Lemma 23.

Let Z𝒩(0,1)Z\sim\mathcal{N}(0,1) be a standard Normal variable. Assume β(1/2,1)\beta\in(1/2,1) and GtG_{t} is a self-similar Gaussian process satisfing 1 and 𝔼[G12]=1\mathbb{E}[G_{1}^{2}]=1. When TT is large enough, there exists a constant Cβ,VC_{\beta,V}^{\prime} such that

supz|(FTHTz)(Zz)|Cβ,VT22β,\sup_{z\in\mathbb{R}}\left|\mathbb{P}\left(\frac{F_{T}^{*}}{H_{T}^{*}}\leq z\right)-\mathbb{P}(Z\leq z)\right|\leq\frac{C_{\beta,V}^{\prime}}{T^{2-2\beta}}, (46)
Proof.

According to Lemma 19, (45f) and (45g), we have

h1,T2K1T22β,h2,T22K1T,\displaystyle\left\|h_{1,T^{*}}\right\|_{\mathfrak{H}}^{2}\leq\frac{K_{1}}{T^{2-2\beta}},\;\;\;\left\|h_{2,T^{*}}\right\|_{\mathfrak{H}^{\otimes 2}}^{2}\leq\frac{K_{1}}{T},

where K1K_{1} is a constant independent of TT. It is easy to see that Lemma 19 and equation (45e) implies 𝔼HTαa.s.\mathbb{E}H_{T}^{*}\to\alpha\;\mathrm{a.s.}. Furthermore, when TT is large enough, there exists constant K2K_{2} such that

|(𝔼HT)2α2|\displaystyle\left|(\mathbb{E}H_{T}^{*})^{2}-\alpha^{2}\right| K2T22β.\displaystyle\leq\frac{K_{2}}{T^{2-2\beta}}. (47)

Since GtG_{t} is self-similar, Lemma 19 implies |f1,T2α2|K3T22β\left|\left\|f_{1,T}^{*}\right\|_{\mathfrak{H}}^{2}-\alpha^{2}\right|\leq\frac{K_{3}}{T^{2-2\beta}}. Also, we can obtain

f2,T22K3T2,f3,T32K3T\displaystyle\left\|f_{2,T}^{*}\right\|_{\mathfrak{H}^{\otimes 2}}^{2}\leq\frac{K_{3}}{T^{2}},\;\;\;\left\|f_{3,T}^{*}\right\|_{\mathfrak{H}^{\otimes 3}}^{2}\leq\frac{K_{3}}{T} (48)

Combining above three formulas, we have

J0,2j=1mj!fj,Tj2K4/T22β,\displaystyle J_{0}^{*,2}-\sum_{j=1}^{m}j!\left\|f_{j,T}^{*}\right\|_{\mathfrak{H}^{\otimes j}}^{2}\leq K_{4}/T^{2-2\beta},

Following from Theorem 16 and Cauchy-Schwarz inequality, we obtain the result. ∎

We next consider the non-zero mean part.

Lemma 24.

Assume β(1/2,1)\beta\in(1/2,1) and GtG_{t} is a self-similar Gaussian process satisfing 1 and 𝔼[G12]=1\mathbb{E}[G_{1}^{2}]=1. Let M1=I0/HTM_{1}=I_{0}^{*}/H_{T}^{*}. When TT is large enough, there exists a constant C1C_{1} independent of TT such that

(|M1|>C1T1/2β/2)=0.\mathbb{P}\Bigl{(}\left|M_{1}\right|>\frac{C_{1}}{T^{1/2-\beta/2}}\Bigr{)}=0.
Proof.

Following from Lemma 23, we have |M1|C|I0|\left|M_{1}\right|\leq C^{\prime}\left|I_{0}^{*}\right|, where CC^{\prime} is a constant independent of TT. Furthermore, Equation (45a) implies that there exists C′′C^{\prime\prime} such that

|I0|C′′T1β,|M1|C1T1βa.s.,\displaystyle\left|I_{0}^{*}\right|\leq\frac{C^{\prime\prime}}{T^{1-\beta}},\;\;\;\left|M_{1}\right|\leq\frac{C_{1}}{T^{1-\beta}}\;\;\;\mathrm{a.s.},

where C1=2CC′′C_{1}=2C^{\prime}\cdot C^{\prime\prime}. Combining the above formulas, we obtain the desired result. ∎

We now prove the formula (11).

Proof of formula (11).

Following from Lemma 10, we have

supz|(kTβ1(μ^LSμ)z)(Zz)|\displaystyle\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\frac{k}{T^{\beta-1}}(\hat{\mu}_{LS}-\mu)\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right| supz|(FTHTz)(Zz)|\displaystyle\leq\sup_{z\in\mathbb{R}}\left|\mathbb{P}\Bigl{(}\frac{F_{T}^{*}}{H_{T}^{*}}\leq z\Bigr{)}-\mathbb{P}(Z\leq z)\right|
+(|M1|>C1T1/2β/2)\displaystyle\ \ \ +\mathbb{P}\Bigl{(}\left|M_{1}\right|>\frac{C_{1}}{T^{1/2-\beta/2}}\Bigr{)}
+12πC1T1/2β/2.\displaystyle\ \ \ +\frac{1}{\sqrt{2\pi}}\frac{C_{1}}{T^{1/2-\beta/2}}.

Combining Lemmas 23 and 24, we obtain the formula (11). ∎

Acknowledgments

We gratefully acknowledge the very valuable suggestions by referees. Y. Chen is supported by National Natural Science Foundation of China (NSFC) with grant No.11961033.

Declarations

  • Availability of data and materials

This manuscript has no associated data.

References

  • Vasicek [1977] Oldrich Vasicek. An equilibrium characterization of the term structure. Journal of Financial Economics, 5(2):177–188, 1977.
  • Wu et al. [2020] Sang Wu, YingHui Dong, WenXin Lv, and GuoJing Wang. Optimal asset allocation for participating contracts with mortality risk under minimum guarantee. Communications in Statistics - Theory and Methods, 49(14):3481–3497, 2020.
  • Han and Li [2021] YueCai Han and Nan Li. Calibrating fractional vasicek model. Communications in Statistics - Theory and Methods, pages 1–15, 2021.
  • Fergusson and Platen [2015] Kevin Fergusson and Eckhard Platen. Application of maximum likelihood estimatation to stochastic short rate models. Annals of Financial Economics, 10(02):1–26, 2015. doi: 10.1142/S2010495215500098. URL http://www.worldscientific.com/doi/10.1142/S2010495215500098.
  • Xiao and Yu [2019] WeiLin Xiao and Jun Yu. Asymptotic theory for rough fractional vasicek models. Economics Letters, 177:26–29, 2019. ISSN 0165-1765. doi: https://doi.org/10.1016/j.econlet.2019.01.020. URL https://www.sciencedirect.com/science/article/pii/S0165176519300266.
  • Hu and Nualart [2010] YaoZhong Hu and David Nualart. Parameter estimation for fractional ornstein–uhlenbeck processes. Statistics & Probability Letters, 80(11-12):1030–1038, 2010.
  • Chen and Zhou [2021] Yong Chen and HongJuan Zhou. Parameter estimation for an ornstein-uhlenbeck process driven by a general gaussian noise. Acta Mathematica Scientia, 41(2):573–595, 2021.
  • Yu [2018] Qian Yu. Statistical inference for vasicek–type model driven by self–similar gaussian processes. Communications in Statistics - Theory and Methods, pages 1–14, 2018. doi: 10.1080/03610926.2018.1543774. URL http://www.tandfonline.com/doi/full/10.1080/03610926.2018.1543774.
  • Xiao et al. [2018] WeiLin Xiao, XiLi Zhang, and Ying Zuo. Least squares estimation for the drift parameters in the sub–fractional vasicek processes. Journal of Statistical Planning and Inference, 197:141–155, 2018. ISSN 0378-3758. doi: https://doi.org/10.1016/j.jspi.2018.01.003. URL https://www.sciencedirect.com/science/article/pii/S0378375818300053.
  • Cai et al. [2022] ChunHao Cai, QingHua Wang, and WeiLin Xiao. Mixed sub-fractional brownian motion and drift estimation of related ornstein–uhlenbeck process. Communications in Mathematics and Statistics, pages 1–27, 2022.
  • Chen et al. [2021] Yong Chen, Ying Li, and XingZhi Pei. Parameter estimation for vasicek model driven by a general gaussian noise. Communications in Statistics - Theory and Methods, pages 1–17, 2021.
  • Nourdin and Peccati [2012] Ivan Nourdin and Giovanni Peccati. Normal Approximations with Malliavin Calculus: From Stein’s Method to Universality, volume 192. Cambridge University Press, Cambridge, 2012. doi: 10.1017/CBO9781139084659.
  • Peccati [2007] G. Peccati. Gaussian approximations of multiple integrals. Electronic Communications in Probability, 12(23):350–364, 2007. doi: 10.1214/ECP.v12-1322. URL http://www.ams.org/mathscinet-getitem?mr=2350573.
  • Nualart [2006] David Nualart. The Malliavin Calculus and Related Topics. Springer Science & Business Media, Berlin, Heidelberg, 2006. URL http://link.springer.com/978-1-4757-2437-0.
  • Jolis [2007] Maria Jolis. On the wiener integral with respect to the fractional brownian motion on an interval. Journal of Mathematical Analysis and Applications, 330(2):1115–1127, 2007. doi: 10.1016/j.jmaa.2006.07.100. URL http://www.sciencedirect.com/science/article/pii/S0022247X06007864.
  • Chang and Rao [1989] Myron N Chang and PV Rao. Berry-esseen bound for the kaplan-meier estimator. Communications in Statistics - Theory and Methods, 18(12):4647–4664, 1989.
  • Kim and Park [2017] Yoon Tae Kim and Hyun Suk Park. Optimal berry–esseen bound for statistical estimations and its application to spde. Journal of Multivariate Analysis, 155:284–304, 2017.
  • Biermé et al. [2012] Hermine Biermé, Aline Bonami, Ivan Nourdin, and Giovanni Peccati. Optimal berry-esseen rates on the wiener space: the barrier of third and fourth cumulants. ALEA : Latin American Journal of Probability and Mathematical Statistics, 9(2):473–500, 2012. URL https://hal.archives-ouvertes.fr/hal-00620384. 29 pages.
  • Brouste and Kleptsyna [2010] Alexandre Brouste and Marina Kleptsyna. Asymptotic properties of mle for partially observed fractional diffusion system. Statistical Inference for Stochastic Processes, 13(1):1–13, 2010.
  • Chen et al. [2020] Yong Chen, NengHui Kuang, and Ying Li. Berry–esséen bound for the parameter estimation of fractional ornstein–uhlenbeck processes. Stochastics and Dynamics, 20(04):1–11, 2020.
  • Lohvinenko et al. [2016] Stanislav Lohvinenko, Kostiantyn Ralchenko, and Olga Zhuchenko. Asymptotic properties of parameter estimators in fractional vasicek model. Lithuanian Journal of Statistics, 55(1):102–111, 2016.
  • Nourdin and Peccati [2009] Ivan Nourdin and Giovanni Peccati. Stein’s method and exact berry–esseen asymptotics for functionals of gaussian fields. The Annals of Probability, 37(6):2231–2261, 2009.
  • Nourdin and Peccati [2015] Ivan Nourdin and Giovanni Peccati. The optimal fourth moment theorem. Proceedings of the American Mathematical Society, 143(7):3123–3133, 2015.