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


Sojourns of Stationary Gaussian Processes over a Random Interval

Krzysztof Dȩbicki Krzysztof Dȩbicki, Mathematical Institute, University of Wrocław, pl. Grunwaldzki 2/4, 50-384 Wrocław, Poland [email protected]  and  Xiaofan Peng Xiaofan Peng, School of Mathematical Sciences, University of Electronic Science and Technology of China, Chengdu 610054, China [email protected]

Abstract: We investigate asymptotics of the tail distribution of sojourn time

0T𝕀(X(t)>u)𝑑t,\int_{0}^{T}\mathbb{I}(X(t)>u)dt,

as uu\to\infty, where XX is a centered stationary Gaussian process and TT is an independent of XX nonnegative random variable. The heaviness of the tail distribution of TT impacts the form of the asymptotics, leading to four scenarios: the case of integrable TT, the case of regularly varying TT with index λ=1\lambda=1 and index λ(0,1)\lambda\in(0,1) and the case of slowly varying tail distribution of TT. The derived findings are illustrated by the analysis of the class of fractional Ornstein-Uhlenbeck processes.

Key Words: exact asymptotics; regularly varying function; sojourn time; stationary Gaussian process.

AMS Classification: Primary 60G15; secondary 60G70

1. Introduction

For given stochastic process X(t),t0X(t),t\geq 0, by

Lu[a,b]:=ab𝕀u(X(t))dt,\displaystyle L_{u}[a,b]:=\int_{a}^{b}\mathbb{I}_{u}\left(X(t)\right)\text{\rm d}t,

with 𝕀u(x):=𝕀(x>u)\mathbb{I}_{u}\left(x\right):=\mathbb{I}\left(x>u\right), we define the sojourn time spent above a fixed level uu by process XX on interval [a,b][a,b]. The interest in analysis of distributional properties of Lu[a,b]L_{u}[a,b] stems both from theoretical questions related to the research on the level sets of stochastic processes and from its importance in applied probability, as e.g. in finance or insurance theory, where Lu[0,T]L_{u}[0,T], T>0T>0 may be interpreted as the total time in ruin up to time TT for the risk process modeled by XX; see e.g. [1, 2].

In the case of XX being a Gaussian process, the asymptotics of the tail distribution of Lu[0,T]L_{u}[0,T], as uu\to\infty, was analyzed extensively in a series of papers by Berman, e.g. [3, 4]; see also the seminal monograph [5] and recent refinements [6, 7].

The aim of this paper is to get the exact asymptotics of tail distribution of Lu[0,T]L_{u}[0,T] for a class of centered stationary Gaussian processes over an independent of XX random time TT. The motivation to consider extremal behaviour of a stochastic process over a random time interval stems from its relevance in such problems as ruin of time-changed risk processes [8, 9], resetting models [10] or hybrid queueing models [11]. We also refer to related problems on extremes of conditionally Gaussian processes and Gaussian processes with random variance [12, 13]. Using the fact that

{Lu[0,T]>0}={supt[0,T]X(t)>u},\mathbb{P}\left\{L_{u}[0,T]>0\right\}=\mathbb{P}\left\{\sup_{t\in[0,T]}X(t)>u\right\},

the findings of this contribution also extend results obtained in [14, 15, 16].

It appears that the form of the derived exact asymptotics strongly depends on the heaviness of the tail distribution of TT, leading to four scenarios: the case of finite 𝔼T{\mathbb{E}}T (scenario D1), the case of TT having regularly varying tail distribution with index λ=1\lambda=1 (scenario D2), λ(0,1)\lambda\in(0,1) (scenario D3) and the case of slowly varying tail distribution of TT (scenario D4); see Section 3.

Brief organisation of the rest of the paper: In Section 2 we formalize the analyzed model and introduce notation. In Section 3 we derive the tail asymptotic behavior of the sojourn time for a class of centered stationary Gaussian processes XX over random interval [0,T][0,T] under introduced in Section 2 scenarios D1-D4, respectively. Section 4 contains some examples illustrating the main findings of this contribution. All the proofs are displayed in Section 5, whereas few technical results are included in Section 6.

2. Notation and model description

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with a.s. continuous trajectories, unit variance function and covariance function rr satisfying

  • A1:

    1r(t)1-r(t) is regularly varying at t=0t=0 with index α(0,2]\alpha\in(0,2];

  • A2:

    r(t)<1r(t)<1 for all t>0t>0;

  • A3:

    limtr(t)log(t)=0\lim_{t\to\infty}r(t)\log(t)=0.

Assumptions A1-A3 cover wide range of investigated in the literature stationary Gaussian processes, where A3 is referred to as Berman’s condition (see, e.g., [5]); see also Section 4.

Let function v()v(\cdot) be such that limuv(u)=\lim_{u\to\infty}v(u)=\infty and

(2.1) limuu2(1r(1/v(u)))=1.\displaystyle\lim_{u\to\infty}u^{2}(1-r(1/v(u)))=1.

By [5], v()v(\cdot) exists and is regularly varying at infinity with index 2/α2/\alpha.

We are interested in the asymptotics of

{Lu[0,T]>x},\mathbb{P}\left\{L_{u}^{*}[0,T]>x\right\},

as uu\to\infty, where

(2.2) Lu[0,T]:=v(u)Lu[0,T]\displaystyle L_{u}^{*}[0,T]:=v(u)L_{u}[0,T]

and TT is an independent of XX nonnegative random variable with distribution function FT()F_{T}(\cdot) which belongs to one of the following distribution classes:

  • D1:

    TT is integrable;

  • D2:

    TT has regularly varying tail distribution with index λ=1\lambda=1;

  • D3:

    TT has regularly varying tail distribution with index λ(0,1)\lambda\in(0,1);

  • D4:

    TT has slowly varying tail distribution.

Define for any x0x\geq 0

(2.3) α(x)=limSS1α(S,x),\displaystyle\mathcal{B}_{\alpha}(x)=\lim_{S\to\infty}S^{-1}\mathcal{B}_{\alpha}(S,x),

with

(2.4) α(S,x)={0S𝕀0(Wα(s)+z)ds>x}ezdz,Wα(t)=2Bα(t)|t|α,\displaystyle\mathcal{B}_{\alpha}(S,x)=\int_{\mathbb{R}}\mathbb{P}\left\{\int_{0}^{S}\mathbb{I}_{0}\left(W_{\alpha}(s)+z\right)\text{\rm d}s>x\right\}e^{-z}\text{\rm d}z,\quad W_{\alpha}(t)=\sqrt{2}B_{\alpha}(t)-\left\lvert t\right\rvert^{\alpha},

where BαB_{\alpha} is a standard fractional Brownian motion (fBm) with Hurst index α/2(0,1]\alpha/2\in(0,1]. By Theorem 2.1 in [6], we know that α(x)\mathcal{B}_{\alpha}(x) is positive and finite for any x0x\geq 0. Let \mathcal{E} be a unit exponential random variable independent of WαW_{\alpha} and set

𝒢α(x)={𝕀0(Wα(s)+)dsx}.\displaystyle\mathcal{G}_{\alpha}(x)=\mathbb{P}\left\{\int_{\mathbb{R}}\mathbb{I}_{0}\left(W_{\alpha}(s)+\mathcal{E}\right)\text{\rm d}s\leq x\right\}.

As shown in [7], 𝒢α\mathcal{G}_{\alpha} is continuous on +\mathbb{R}^{+}, and thus by Remark 2.2 ii) in [6]

α(x)=x1yd𝒢α(y)\mathcal{B}_{\alpha}(x)=\int_{x}^{\infty}\frac{1}{y}\text{\rm d}\mathcal{G}_{\alpha}(y)

holds for all x+x\in\mathbb{R}^{+}. We note that α(0)\mathcal{B}_{\alpha}(0) is equal to the classical Pickands constant; see e.g. [17] or Section 10 in [5]. Let

(2.5) m(u)=(α(0)v(u)Ψ(u))1,\displaystyle m(u)=\big{(}\mathcal{B}_{\alpha}(0)v(u)\Psi(u)\big{)}^{-1},

where Ψ(u)\Psi(u) is the survival function of an N(0,1)N(0,1) random variable. Then, by Theorem 10.5.1 in [5],

(2.6) {supt[0,1]X(t)>u}m1(u),u.\displaystyle\mathbb{P}\left\{\sup_{t\in[0,1]}X(t)>u\right\}\sim m^{-1}(u),\quad u\to\infty.

In our notation \sim stands for asymptotic equivalence of two functions as the argument tends to 0 or to \infty respectively.

3. Main results

In this section we find the exact asymptotics of

(3.1) {Lu[0,T]>x}\displaystyle\mathbb{P}\left\{L_{u}^{*}[0,T]>x\right\}

as uu\to\infty, under scenarios D1-D4, respectively. All the proofs are postponed to Section 5.

3.1. Scenario D1

We begin with the case when TT is integrable. It appears that under this scenario the main contribution to the asymptotics of (3.1) comes from Gaussian process XX, whereas TT contributes only by its average behavior.

Theorem 3.1.

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with unit variance and covariance function satisfying A1-A2. Suppose that TT is an independent of XX nonnegative random variable that satisfies D1. Then for any x0x\geq 0

(3.2) {Lu[0,T]>x}α(x)𝔼{T}v(u)Ψ(u),u.\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathcal{B}_{\alpha}(x)\mathbb{E}\left\{T\right\}v(u)\Psi(u),\quad u\to\infty.

We can rewrite the result in Theorem 3.1 as

{Lu[0,T]>x}𝔼{T}{Lu[0,1]>x},u.\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathbb{E}\left\{T\right\}\mathbb{P}\left\{L_{u}^{\ast}[0,1]>x\right\},\quad u\to\infty.

3.2. Scenario D2

Under this scenario the asymptotics of (3.1) is similar to the one obtained for case D1 with the exception that TT contributes to (3.1) by its integrated tail distribution rather than by its mean.

Theorem 3.2.

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with unit variance and covariance function satisfying A1-A3. Suppose that TT is an independent of XX nonnegative random variable that satisfies D2. Then for any x0x\geq 0

(3.3) {Lu[0,T]>x}α(x)l(m(u))v(u)Ψ(u),u,\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathcal{B}_{\alpha}(x)l(m(u))v(u)\Psi(u),\quad u\to\infty,

where l(u)=0u{T>t}dtl(u)=\int_{0}^{u}\mathbb{P}\left\{T>t\right\}\text{\rm d}t.

Remark 3.3.

We note that if TT satisfies D2 and is integrable, then (3.3) coincides with (3.2).

3.3. Scenario D3

This scenario leads to the asymptotics of (3.1) which depends only of the heaviness of the tail distribution of TT.

The following continuous distribution function

(3.4) α(x):=α1(0)0x1yd𝒢α(x),x0\displaystyle\mathcal{F}_{\alpha}(x):=\mathcal{B}_{\alpha}^{-1}(0)\int_{0}^{x}\frac{1}{y}\text{\rm d}\mathcal{G}_{\alpha}(x),\quad x\geq 0

plays an important role in further analysis. αk¯(x)\overline{\mathcal{F}_{\alpha}^{*k}}(x) denotes the tail distribution of the kk-th convolution of α\mathcal{F}_{\alpha} at x0x\geq 0.

Theorem 3.4.

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with unit variance and covariance function satisfying A1-A3. Suppose that TT is an independent of XX nonnegative random variable that satisfies D3. Then for any x0x\geq 0

(3.5) {Lu[0,T]>x}λk=1Γ(kλ)k!αk¯(x){T>m(u)},u.\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\lambda\sum_{k=1}^{\infty}\frac{\Gamma(k-\lambda)}{k!}\overline{\mathcal{F}_{\alpha}^{*k}}(x)\mathbb{P}\left\{T>m(u)\right\},\quad u\to\infty.
Remark 3.5.

Taking x=0x=0 in (3.5) and using

λk=1Γ(kλ)k!=λk=11k!0lkλ1eldl=λ0(1el)lλ1dl=Γ(1λ),\lambda\sum_{k=1}^{\infty}\frac{\Gamma(k-\lambda)}{k!}=\lambda\sum_{k=1}^{\infty}\frac{1}{k!}\int_{0}^{\infty}l^{k-\lambda-1}e^{-l}\text{\rm d}l=\lambda\int_{0}^{\infty}(1-e^{-l})l^{-\lambda-1}\text{\rm d}l=\Gamma(1-\lambda),

we recover Theorem 3.2 in [16].

3.4. Scenario D4

Suppose now that TT has slowly varying tail distribution at \infty. As shown in the following theorem, similarly to case D3, the asymptotics of (3.1) depends only on the asymptotic behavior of the tail distribution of TT but in contrast to scenario D3 doesn’t depend on xx.

Theorem 3.6.

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with unit variance and covariance function satisfying A1-A3. Suppose that TT is an independent of XX nonnegative random variable that satisfies D4. Then for any x0x\geq 0

{Lu[0,T]>x}{T>m(u)},u.\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathbb{P}\left\{T>m(u)\right\},\quad u\to\infty.

4. Examples

In this section we illustrate the results derived in Section 3 by two classes of stationary Gaussian processes: fractional Ornstein-Uhlenbeck processes and increments of fractional Brownian motions.

4.1. Fractional Ornstein-Uhlenbeck processes

Suppose that XX is a centered stationary Gaussian process with covariance r(t)=etα,t0r(t)=e^{-t^{\alpha}},t\geq 0, for α(0,2]\alpha\in(0,2]. We call XX a fractional Urnstein-Uhlenbeck process with index α\alpha. If α=1\alpha=1, then XX is the classical Ornstein-Uhlenbeck process.

It is straightforward to check that A1-A3 are satisfied. Thus, the following proposition holds due to Theorems 3.1-3.6.

Proposition 4.1.

Suppose that XX is a fractional Ornstein-Uhlenbeck process with index α(0,2]\alpha\in(0,2], and TT is an independent of XX nonnegative random variable. Then for any x0x\geq 0, as uu\to\infty,
(i) If TT\in
D1, then {Lu[0,T]>x}α(x)𝔼{T}(2π)1/2u2/α1eu2/2\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathcal{B}_{\alpha}(x)\mathbb{E}\left\{T\right\}(2\pi)^{-1/2}u^{2/\alpha-1}e^{-u^{2}/2}.
(ii) If TT\in
D2, then {Lu[0,T]>x}α(x)(2π)1/2u2/α1eu2/202πα1(0)u12/αeu2/2{T>t}dt.\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathcal{B}_{\alpha}(x)(2\pi)^{-1/2}u^{2/\alpha-1}e^{-u^{2}/2}\int_{0}^{\sqrt{2\pi}\mathcal{B}_{\alpha}^{-1}(0)u^{1-2/\alpha}e^{u^{2}/2}}\mathbb{P}\left\{T>t\right\}\text{\rm d}t.
(iii) If TT\in
D3, then {Lu[0,T]>x}λk=1Γ(kλ)k!αk¯(x){T>2πα1(0)u12/αeu2/2}\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\lambda\sum_{k=1}^{\infty}\frac{\Gamma(k-\lambda)}{k!}\overline{\mathcal{F}_{\alpha}^{*k}}(x)\mathbb{P}\left\{T>\sqrt{2\pi}\mathcal{B}_{\alpha}^{-1}(0)u^{1-2/\alpha}e^{u^{2}/2}\right\}.
(iv) If TT\in
D4, then {Lu[0,T]>x}{T>2πα1(0)u12/αeu2/2}\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathbb{P}\left\{T>\sqrt{2\pi}\mathcal{B}_{\alpha}^{-1}(0)u^{1-2/\alpha}e^{u^{2}/2}\right\}.

4.2. Increments of fractional Brownian motion

For a standard fBm Bα(t),t0B_{\alpha}(t),t\geq 0 with Hurst index α/2(0,1)\alpha/2\in(0,1) and a>0a>0, define

Xα,a(t):=Bα(t+a)Bα(t)aα/2,t0.\displaystyle X_{\alpha,a}(t):=\frac{B_{\alpha}(t+a)-B_{\alpha}(t)}{a^{\alpha/2}},\quad t\geq 0.

One can check that Xα,aX_{\alpha,a} is a centered stationary Gaussian process with unit variance and covariance function

r(t)=(a+t)α+|at|α2tα2aα,t0.r(t)=\frac{(a+t)^{\alpha}+\left\lvert a-t\right\rvert^{\alpha}-2t^{\alpha}}{2a^{\alpha}},\quad t\geq 0.

and 1r(t)aαtα,1-r(t)\sim a^{-\alpha}t^{\alpha}, t0,t\to 0,which verifies assumption A1. Similarly for t>at>a

|r(t)|α|1α|(ta)α22aα2,\displaystyle\left\lvert r(t)\right\rvert\leq\frac{\alpha\left\lvert 1-\alpha\right\rvert(t-a)^{\alpha-2}}{2a^{\alpha-2}},

which confirms assumption A3. Thus the following proposition holds.

Proposition 4.2.

Suppose that Xα,a(t),t0X_{\alpha,a}(t),t\geq 0 with α(0,2)\alpha\in(0,2), a>0a>0 is independent of a nonnegative random variable TT. Then for any x0x\geq 0, as uu\to\infty,
(i) If TT\in
D1, then {Lu[0,T]>x}α(x)𝔼{T}(2π)1/2a1u2/α1eu2/2\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathcal{B}_{\alpha}(x)\mathbb{E}\left\{T\right\}(2\pi)^{-1/2}a^{-1}u^{2/\alpha-1}e^{-u^{2}/2}.
(ii) If TT\in
D2, then

{Lu[0,T]>x}α(x)(2π)1/2a1u2/α1eu2/202πaα1(0)u12/αeu2/2{T>t}dt.\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathcal{B}_{\alpha}(x)(2\pi)^{-1/2}a^{-1}u^{2/\alpha-1}e^{-u^{2}/2}\int_{0}^{\sqrt{2\pi}a\mathcal{B}_{\alpha}^{-1}(0)u^{1-2/\alpha}e^{u^{2}/2}}\mathbb{P}\left\{T>t\right\}\text{\rm d}t.

(iii) If TT\in D3, then {Lu[0,T]>x}λk=1Γ(kλ)k!αk¯(x){T>2πaα1(0)u12/αeu2/2};\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\lambda\sum_{k=1}^{\infty}\frac{\Gamma(k-\lambda)}{k!}\overline{\mathcal{F}_{\alpha}^{*k}}(x)\mathbb{P}\left\{T>\sqrt{2\pi}a\mathcal{B}_{\alpha}^{-1}(0)u^{1-2/\alpha}e^{u^{2}/2}\right\};
(iv) If TT\in
D4, then {Lu[0,T]>x}{T>2πaα1(0)u12/αeu2/2}\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}\sim\mathbb{P}\left\{T>\sqrt{2\pi}a\mathcal{B}_{\alpha}^{-1}(0)u^{1-2/\alpha}e^{u^{2}/2}\right\}.

5. Proofs

In this section we give detailed proofs of all the theorems presented in Section 3. We first give a simple extension of Theorem 7.4.1 of [5].

Lemma 5.1.

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with unit variance and covariance function satisfying A1 and A3. If LuL_{u}^{\ast}, m(u)m(u) and α\mathcal{F}_{\alpha} are defined in (2.2), (2.5) and (3.4), respectively, then for any s0s\geq 0 and 0<l0<l1<0<l_{0}<l_{1}<\infty we have

(5.1) limusupτ[l0,l1]|𝔼{esLu[0,τm(u)]}eτ0(1esx)dα(x)|=0.\displaystyle\lim_{u\to\infty}\sup_{\tau\in[l_{0},l_{1}]}\left\lvert\mathbb{E}\left\{e^{-sL_{u}^{\ast}[0,\tau m(u)]}\right\}-e^{-\tau\int_{0}^{\infty}(1-e^{-sx})\text{\rm d}\mathcal{F}_{\alpha}(x)}\right\rvert=0.

Proof of Lemma 5.1 For any τ>0\tau>0, the point convergence follows from Berman’s proof of Theorem 7.4.1 in [5]. The uniformity of the convergence on [l0,l1][l_{0},l_{1}] follows by monotonicity of 𝔼{esLu[0,τm(u)]}\mathbb{E}\left\{e^{-sL_{u}^{\ast}[0,\tau m(u)]}\right\} and by continuity of eτ0(1esx)dα(x)e^{-\tau\int_{0}^{\infty}(1-e^{-sx})\text{\rm d}\mathcal{F}_{\alpha}(x)} as function of τ\tau. \Box

Define a compound Poisson process

(5.2) Y(t)=i=1N(t)ξi,\displaystyle Y(t)=\sum_{i=1}^{N(t)}\xi_{i},

where {N(t):t0}\{N(t):t\geq 0\} is a Poisson process with unit intensity, and {ξi:i1}\{\xi_{i}:i\geq 1\} are independent and identically distributed random variables, with distribution function α\mathcal{F}_{\alpha}, which are also independent of NN. The following corollary of Lemma 5.1 will play an important role in the proof of Theorem 3.4.

Corollary 5.1.

If XX is the Gaussian process given as in Lemma 5.1 and YY is defined in (5.2), then for any x0x\geq 0 and 0<l0<l1<0<l_{0}<l_{1}<\infty we have

(5.3) limusupl[l0,l1]|{Lu[0,lm(u)]>x}{Y(l)>x}|=0.\displaystyle\lim_{u\to\infty}\sup_{l\in[l_{0},l_{1}]}\left\lvert\mathbb{P}\left\{L_{u}^{\ast}[0,lm(u)]>x\right\}-\mathbb{P}\left\{Y(l)>x\right\}\right\rvert=0.

Proof of Corollary 5.1 For arbitrary l>0l>0, by (5.1),

{Lu[0,lm(u)]>x}{Y(l)>x},u\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,lm(u)]>x\right\}\rightarrow\mathbb{P}\left\{Y(l)>x\right\},\quad u\to\infty

holds for any x>0x>0. Further,

|{Y(l1)>x}{Y(l0)>x}|{Y(|l1l0|)>0}=1e|l1l0||l1l0|,\displaystyle\left\lvert\mathbb{P}\left\{Y(l_{1})>x\right\}-\mathbb{P}\left\{Y(l_{0})>x\right\}\right\rvert\leq\mathbb{P}\left\{Y(\left\lvert l_{1}-l_{0}\right\rvert)>0\right\}=1-e^{-\left\lvert l_{1}-l_{0}\right\rvert}\leq\left\lvert l_{1}-l_{0}\right\rvert,

which implies that for any x>0x>0, {Y(l)>x}\mathbb{P}\left\{Y(l)>x\right\} is continuous in ll. Finally, the uniform convergence follows by the same argument as stated in Lemma 5.1. For x=0x=0 in (5.3), we refer to Lemma 4.3 in [16]. This completers the proof. \Box

5.1. Proof of Theorem 3.1

By (2.6), for arbitrary ε>0\varepsilon>0, there exists large enough uu such that

{supt[0,1]X(t)>u}<(1+ε)m1(u),\mathbb{P}\left\{\sup_{t\in[0,1]}X(t)>u\right\}<(1+\varepsilon)m^{-1}(u),

which together with the stationarity of process XX implies that for any x0x\geq 0 and t>0t>0

{Lu[0,t]>x}v(u)Ψ(u)\displaystyle\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,t]>x\right\}}{v(u)\Psi(u)} \displaystyle\leq {sups[0,t]X(s)>u}v(u)Ψ(u)\displaystyle\frac{\mathbb{P}\left\{\sup_{s\in[0,t]}X(s)>u\right\}}{v(u)\Psi(u)}
\displaystyle\leq (t+1){sups[0,1]X(s)>u}v(u)Ψ(u)\displaystyle(t+1)\frac{\mathbb{P}\left\{\sup_{s\in[0,1]}X(s)>u\right\}}{v(u)\Psi(u)}
\displaystyle\leq (t+1)(1+ε)α(0).\displaystyle(t+1)(1+\varepsilon)\mathcal{B}_{\alpha}(0).

Consequently, for nonnegative random variable TT with distribution function FTF_{T} satisfying D1, by dominated convergence theorem and Remark 2.2 i) in [6] we have

limu{Lu[0,T]>x}v(u)Ψ(u)\displaystyle\lim_{u\to\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}}{v(u)\Psi(u)} =\displaystyle= limu0{Lu[0,t]>x}v(u)Ψ(u)dFT(t)\displaystyle\lim_{u\to\infty}\int_{0}^{\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,t]>x\right\}}{v(u)\Psi(u)}\text{\rm d}F_{T}(t)
=\displaystyle= α(x)0tdFT(t)\displaystyle\mathcal{B}_{\alpha}(x)\int_{0}^{\infty}t\text{\rm d}F_{T}(t)
=\displaystyle= α(x)𝔼{T}.\displaystyle\mathcal{B}_{\alpha}(x)\mathbb{E}\left\{T\right\}.

This completes the proof. \Box

5.2. Proof of Theorem 3.2

Let A(u)A(u) satisfy

limuA(u)v(u)=andlimuA(u)=0.\displaystyle\lim_{u\to\infty}A(u)v(u)=\infty\quad\textrm{and}\quad\lim_{u\to\infty}A(u)=0.

By Corollary 6.1, for any x0x\geq 0 and arbitrary ε(0,1)\varepsilon\in(0,1), there exist δ>0\delta>0 and u0u_{0} such that

inft[A(u),δm(u)]{Lu[0,t]>x}tα(x)v(u)Ψ(u)12ε,u>u0\displaystyle\inf_{t\in[A(u),\delta m(u)]}\frac{\mathbb{P}\left\{L_{u}^{*}[0,t]>x\right\}}{t\mathcal{B}_{\alpha}(x)v(u)\Psi(u)}\geq 1-2\varepsilon,\quad u>u_{0}

and

supt[A(u),δm(u)]{Lu[0,t]>x}tα(x)v(u)Ψ(u)1+2ε,u>u0.\displaystyle\sup_{t\in[A(u),\delta m(u)]}\frac{\mathbb{P}\left\{L_{u}^{*}[0,t]>x\right\}}{t\mathcal{B}_{\alpha}(x)v(u)\Psi(u)}\leq 1+2\varepsilon,\quad u>u_{0}.

Therefore,

lim infu{Lu[0,T]>x}α(x)l(m(u))v(u)Ψ(u)\displaystyle\liminf_{u\to\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}}{\mathcal{B}_{\alpha}(x)l(m(u))v(u)\Psi(u)} \displaystyle\geq lim infuA(u)δm(u){Lu[0,t]>x}dFT(t)α(x)l(m(u))v(u)Ψ(u)\displaystyle\liminf_{u\to\infty}\frac{\int_{A(u)}^{\delta m(u)}\mathbb{P}\left\{L_{u}^{\ast}[0,t]>x\right\}\text{\rm d}F_{T}(t)}{\mathcal{B}_{\alpha}(x)l(m(u))v(u)\Psi(u)}
\displaystyle\geq (12ε)lim infuA(u)δm(u)tdFT(t)l(m(u))\displaystyle(1-2\varepsilon)\liminf_{u\to\infty}\frac{\int_{A(u)}^{\delta m(u)}t\text{\rm d}F_{T}(t)}{l(m(u))}
=\displaystyle= (12ε)lim infu0δm(u)tdFT(t)l(m(u))\displaystyle(1-2\varepsilon)\liminf_{u\to\infty}\frac{\int_{0}^{\delta m(u)}t\text{\rm d}F_{T}(t)}{l(m(u))}
=\displaystyle= (12ε)lim infu0δm(u){T>t}dtδm(u){T>δm(u)}l(m(u))\displaystyle(1-2\varepsilon)\liminf_{u\to\infty}\frac{\int_{0}^{\delta m(u)}\mathbb{P}\left\{T>t\right\}\text{\rm d}t-\delta m(u)\mathbb{P}\left\{T>\delta m(u)\right\}}{l(m(u))}
=\displaystyle= (12ε),\displaystyle(1-2\varepsilon),

where the last inequality follows by Proposition 1.5.9a in [18] such that l(u)l(u) is slowly varying at \infty and limuu{T>u}/l(u)=0\lim_{u\to\infty}u\mathbb{P}\left\{T>u\right\}/l(u)=0.

Similarly,

lim supu{Lu[0,T]>x}α(x)l(m(u))v(u)Ψ(u)\displaystyle\quad\limsup_{u\to\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}}{\mathcal{B}_{\alpha}(x)l(m(u))v(u)\Psi(u)}
lim supu{Lu[0,A(u)]>x}{TA(u)}+A(u)δm(u){Lu[0,t]>x}dFT(t)+{T>δm(u)}α(x)l(m(u))v(u)Ψ(u)\displaystyle\leq\limsup_{u\to\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,A(u)]>x\right\}\mathbb{P}\left\{T\leq A(u)\right\}+\int_{A(u)}^{\delta m(u)}\mathbb{P}\left\{L_{u}^{\ast}[0,t]>x\right\}\text{\rm d}F_{T}(t)+\mathbb{P}\left\{T>\delta m(u)\right\}}{\mathcal{B}_{\alpha}(x)l(m(u))v(u)\Psi(u)}
lim supuA(u){TA(u)}l(m(u))+(1+2ε)lim supuA(u)δm(u)tdFT(t)l(m(u))\displaystyle\leq\limsup_{u\to\infty}\frac{A(u)\mathbb{P}\left\{T\leq A(u)\right\}}{l(m(u))}+(1+2\varepsilon)\limsup_{u\to\infty}\frac{\int_{A(u)}^{\delta m(u)}t\text{\rm d}F_{T}(t)}{l(m(u))}
=(1+2ε),\displaystyle=(1+2\varepsilon),

where the last inequality follows by (6.20) and the same reasons as above. Since ε\varepsilon is arbitrary, letting ε0\varepsilon\to 0, we complete the proof. \Box

5.3. Proof of Theorem 3.4

First, note that by Raabe’s Test, the series in (3.5) converges for λ(0,1)\lambda\in(0,1). Then by integration by parts, for any x0x\geq 0 we have

(5.4) 0lλd{Y(l)>x}\displaystyle\int_{0}^{\infty}l^{-\lambda}\text{\rm d}\mathbb{P}\left\{Y(l)>x\right\} =\displaystyle= 0{Y(l)>x}λlλ1dlliml0lλ{Y(l)>x}\displaystyle\int_{0}^{\infty}\mathbb{P}\left\{Y(l)>x\right\}\lambda l^{-\lambda-1}\text{\rm d}l-\lim_{l\to 0}l^{-\lambda}\mathbb{P}\left\{Y(l)>x\right\}
=\displaystyle= λk=1αk¯(x)1k!0lkλ1eldlliml0lλ{Y(l)>x}\displaystyle\lambda\sum_{k=1}^{\infty}\overline{\mathcal{F}_{\alpha}^{*k}}(x)\frac{1}{k!}\int_{0}^{\infty}l^{k-\lambda-1}e^{-l}\text{\rm d}l-\lim_{l\to 0}l^{-\lambda}\mathbb{P}\left\{Y(l)>x\right\}
=\displaystyle= λk=1Γ(kλ)k!αk¯(x)<,\displaystyle\lambda\sum_{k=1}^{\infty}\frac{\Gamma(k-\lambda)}{k!}\overline{\mathcal{F}_{\alpha}^{*k}}(x)<\infty,

where the last equality, recalling that λ(0,1)\lambda\in(0,1), follows by

(5.5) liml0lλ{Y(l)>x}liml0lλ(1el)=0.\displaystyle\lim_{l\to 0}l^{-\lambda}\mathbb{P}\left\{Y(l)>x\right\}\leq\lim_{l\to 0}l^{-\lambda}(1-e^{-l})=0.

Next, by a similar argument as used in the proof of Theorem 3.2 in [16], for any 0<l0<l1<0<l_{0}<l_{1}<\infty we have

{Lu[0,T]>x}\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\} =\displaystyle= (0l0m(u)+l0m(u)l1m(u)+l1m(u)){Lu[0,l]>x}dFT(l)\displaystyle\Big{(}\int_{0}^{l_{0}m(u)}+\int_{l_{0}m(u)}^{l_{1}m(u)}+\int_{l_{1}m(u)}^{\infty}\Big{)}\mathbb{P}\left\{L_{u}^{\ast}[0,l]>x\right\}\text{\rm d}F_{T}(l)
=\displaystyle= I1+I2+I3,\displaystyle I_{1}+I_{2}+I_{3},

where

lim supuI1{T>m(u)}lim supu0l0m(u){sups[0,l]X(s)>u}dFT(l){T>m(u)}λ1λl01λ\displaystyle\limsup_{u\to\infty}\frac{I_{1}}{\mathbb{P}\left\{T>m(u)\right\}}\leq\limsup_{u\to\infty}\frac{\int_{0}^{l_{0}m(u)}\mathbb{P}\left\{\sup_{s\in[0,l]}X(s)>u\right\}\text{\rm d}F_{T}(l)}{\mathbb{P}\left\{T>m(u)\right\}}\leq\frac{\lambda}{1-\lambda}l_{0}^{1-\lambda}

and

lim supuI3{T>m(u)}lim supu{T>l1m(u)}{T>m(u)}=l1λ.\displaystyle\limsup_{u\to\infty}\frac{I_{3}}{\mathbb{P}\left\{T>m(u)\right\}}\leq\limsup_{u\to\infty}\frac{\mathbb{P}\left\{T>l_{1}m(u)\right\}}{\mathbb{P}\left\{T>m(u)\right\}}=l_{1}^{-\lambda}.

Further, in view of Corollary 5.1, for any given x0x\geq 0 and arbitrary ε>0\varepsilon>0, we have the following upper bound

I2\displaystyle I_{2} =\displaystyle= l0l1{Lu[0,lm(u)]>x}dFT(lm(u))\displaystyle\int_{l_{0}}^{l_{1}}\mathbb{P}\left\{L_{u}^{*}[0,lm(u)]>x\right\}\text{\rm d}F_{T}(lm(u))
\displaystyle\leq (1+ε)l0l1{Y(l)>x}dFT(lm(u))\displaystyle(1+\varepsilon)\int_{l_{0}}^{l_{1}}\mathbb{P}\left\{Y(l)>x\right\}\text{\rm d}F_{T}(lm(u))
=\displaystyle= (1+ε)(l0l1{T>lm(u)}d{Y(l)>x}\displaystyle(1+\varepsilon)\Big{(}\int_{l_{0}}^{l_{1}}\mathbb{P}\left\{T>lm(u)\right\}\text{\rm d}\mathbb{P}\left\{Y(l)>x\right\}
{Y(l1)>x}{T>l1m(u)}+{Y(l0)>x}{T>l0m(u)}),\displaystyle\qquad\qquad-\mathbb{P}\left\{Y(l_{1})>x\right\}\mathbb{P}\left\{T>l_{1}m(u)\right\}+\mathbb{P}\left\{Y(l_{0})>x\right\}\mathbb{P}\left\{T>l_{0}m(u)\right\}\Big{)},

which holds for uu large enough. By Potter’s Theorem (see, e.g., [18][Theorem 1.5.6]), for sufficiently large uu there exists some constant C>1C>1 such that

{T>lm(u)}{T>m(u)}Cl2λ\frac{\mathbb{P}\left\{T>lm(u)\right\}}{\mathbb{P}\left\{T>m(u)\right\}}\leq Cl^{-2\lambda}

holds for all l[l0,l1]l\in[l_{0},l_{1}], and thus by dominated convergence theorem

lim supuI2{T>m(u)}(1+ε)(l0l1lλd{Y(l)>x}{Y(l1)>x}l1λ+{Y(l0)>x}l0λ).\displaystyle\limsup_{u\to\infty}\frac{I_{2}}{\mathbb{P}\left\{T>m(u)\right\}}\leq(1+\varepsilon)\Big{(}\int_{l_{0}}^{l_{1}}l^{-\lambda}\text{\rm d}\mathbb{P}\left\{Y(l)>x\right\}-\mathbb{P}\left\{Y(l_{1})>x\right\}l_{1}^{-\lambda}+\mathbb{P}\left\{Y(l_{0})>x\right\}l_{0}^{-\lambda}\Big{)}.

Similarly, we have the lower bound

lim infuI2{T>m(u)}(1ε)(l0l1lλd{Y(l)>x}{Y(l1)>x}l1λ+{Y(l0)>x}l0λ).\displaystyle\liminf_{u\to\infty}\frac{I_{2}}{\mathbb{P}\left\{T>m(u)\right\}}\geq(1-\varepsilon)\Big{(}\int_{l_{0}}^{l_{1}}l^{-\lambda}\text{\rm d}\mathbb{P}\left\{Y(l)>x\right\}-\mathbb{P}\left\{Y(l_{1})>x\right\}l_{1}^{-\lambda}+\mathbb{P}\left\{Y(l_{0})>x\right\}l_{0}^{-\lambda}\Big{)}.

Finally, letting l00l_{0}\to 0 and l1l_{1}\to\infty in the above bounds, using (5.4) and (5.5), and then by the fact that ε>0\varepsilon>0 was arbitrary, we complete the proof. \Box

5.4. Proof of Theorem 3.6

According to Remark 3.3 in [16], we know that

lim supu{Lu[0,T]>x}{T>m(u)}lim supu{sups[0,T]X(s)>u}{T>m(u)}1.\displaystyle\limsup_{u\to\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}}{\mathbb{P}\left\{T>m(u)\right\}}\leq\limsup_{u\to\infty}\frac{\mathbb{P}\left\{\sup_{s\in[0,T]}X(s)>u\right\}}{\mathbb{P}\left\{T>m(u)\right\}}\leq 1.

Further, by Corollary 5.1, for arbitrary l>0l>0

{Lu[0,lm(u)]>x}{Y(l)>x}\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,lm(u)]>x\right\}\rightarrow\mathbb{P}\left\{Y(l)>x\right\}

holds for any x0x\geq 0 as uu\to\infty. Thus, for TT with slowly varying tail distribution we get

lim infu{Lu[0,T]>x}{T>m(u)}\displaystyle\liminf_{u\to\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,T]>x\right\}}{\mathbb{P}\left\{T>m(u)\right\}} \displaystyle\geq lim infu{Lu[0,lm(u)]>x}{T>lm(u)}{T>m(u)}\displaystyle\liminf_{u\to\infty}\frac{\mathbb{P}\left\{L_{u}^{\ast}[0,lm(u)]>x\right\}\mathbb{P}\left\{T>lm(u)\right\}}{\mathbb{P}\left\{T>m(u)\right\}}
=\displaystyle= {Y(l)>x},\displaystyle\mathbb{P}\left\{Y(l)>x\right\},

which converges to 11 as ll\to\infty, since by the strong law of large numbers Y(l)/lα1(0)>0Y(l)/l\to\mathcal{B}_{\alpha}^{-1}(0)>0. This completes the proof. \Box

6. Appendix

Hereafter, Ci,iC_{i},i\in\mathbb{N} are positive constants which may be different from line to line. All vectors are column vectors unless otherwise specified. As long as it doesn’t cause confusion we use 𝟎\boldsymbol{0} to denote the 2×12\times 1 column vector or the 2×22\times 2 matrix whose entries are all 0’s. For a given vector (matrix) 𝐐\mathbf{Q}, let |𝐐||\mathbf{Q}| denote vector (matrix) with entries equal to absolute value of respective entries of 𝐐\mathbf{Q}.

Lemma 6.1.

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with unit variance and covariance function satisfying A1 and A3. If v(u)v(u), α(S,x)\mathcal{B}_{\alpha}(S,x) and m(u)m(u) are defined in (2.1), (2.4) and (2.5), respectively, then for any A(u)>1A(u)>1 satisfying

(6.1) lim supuu2logA(u)<\displaystyle\limsup_{u\to\infty}\frac{u^{2}}{\log A(u)}<\infty

we have

limusupdA(u)|{sups1[0,S]X(s1/v(u))>u,sups2[0,S]X(d+s2/v(u))>u}Ψ2(u)α2(S,0)|=0.\displaystyle\ \ \ \ \ \lim_{u\to\infty}\sup_{d\geq A(u)}\left\lvert\frac{\mathbb{P}\left\{\sup_{s_{1}\in[0,S]}X(s_{1}/v(u))>u,\sup_{s_{2}\in[0,S]}X(d+s_{2}/v(u))>u\right\}}{\Psi^{2}(u)}-\mathcal{B}_{\alpha}^{2}(S,0)\right\rvert=0.

Proof of Lemma 6.1 We borrow the argument used in the proof of Theorem 5.1 in [6]. First, for notational simplicity we define

ξu,d(𝒔)=(X(s1/v(u)),X(d+s2/v(u)))T,𝒔=(s1,s2)𝑫=[0,S]2,\displaystyle\xi_{u,d}(\boldsymbol{s})=(X(s_{1}/v(u)),X(d+s_{2}/v(u)))^{T},\quad\boldsymbol{s}=(s_{1},s_{2})\in\boldsymbol{D}=[0,S]^{2},

and denote by Ru,d(𝒔,𝒕)R_{u,d}(\boldsymbol{s},\boldsymbol{t}) the covariance matrix function of ξu,d\xi_{u,d}, i.e.,

Ru,d(𝒔,𝒕)\displaystyle R_{u,d}(\boldsymbol{s},\boldsymbol{t}) =\displaystyle= Cov(ξu,d(𝒔),ξu,d(𝒕))\displaystyle\mathrm{Cov}(\xi_{u,d}(\boldsymbol{s}),\xi_{u,d}(\boldsymbol{t}))
=\displaystyle= 𝔼{ξu,d(𝒔)ξu,d(𝒕)T}\displaystyle\mathbb{E}\left\{\xi_{u,d}(\boldsymbol{s})\xi_{u,d}(\boldsymbol{t})^{T}\right\}
=\displaystyle= (r(|t1s1|v(u))r(d+t2s1v(u))r(d+s2t1v(u))r(|t2s2|v(u)))𝒔,𝒕𝑫.\displaystyle\begin{pmatrix}r(\frac{|t_{1}-s_{1}|}{v(u)})&r(d+\frac{t_{2}-s_{1}}{v(u)})\\ r(d+\frac{s_{2}-t_{1}}{v(u)})&r(\frac{|t_{2}-s_{2}|}{v(u)})\end{pmatrix}\quad\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}.

Then, conditioning on ξu,d(𝟎)\xi_{u,d}(\boldsymbol{0}) we have

{𝒔𝑫ξu,d(𝒔)>𝒖}=2{𝒔𝑫ξu,d(𝒔)>𝒖|ξu,d(𝟎)=𝒚}ϕ(y1,y2;r(d))dy1dy2,\displaystyle\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}\xi_{u,d}(\boldsymbol{s})>\boldsymbol{u}\right\}=\iint_{\mathbb{R}^{2}}\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}\xi_{u,d}(\boldsymbol{s})>\boldsymbol{u}|\xi_{u,d}(\boldsymbol{0})=\boldsymbol{y}\right\}\phi(y_{1},y_{2};r(d))\text{\rm d}y_{1}\text{\rm d}y_{2},

where 𝒚=(y1,y2)T\boldsymbol{y}=(y_{1},y_{2})^{T} and ϕ(y1,y2;r(d))\phi(y_{1},y_{2};r(d)) is the density function of bivariate normal random variable ξu,d(𝟎)\xi_{u,d}(\boldsymbol{0}). By the change of variables 𝒚=𝒖+𝒛/u\boldsymbol{y}=\boldsymbol{u}+\boldsymbol{z}/u and using properties of conditional distribution of normal random variable (see e.g., Chapter 2.2 in [5]), we get

{𝒔𝑫ξu,d(𝒔)>𝒖}\displaystyle\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}\xi_{u,d}(\boldsymbol{s})>\boldsymbol{u}\right\} =\displaystyle= eu22πu22{𝒔𝑫χu,d(𝒔)θu,d(𝒔,𝒛)>𝟎}fu,d(𝒛)dz1dz2\displaystyle\frac{e^{-u^{2}}}{2\pi u^{2}}\iint_{\mathbb{R}^{2}}\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}\chi_{u,d}(\boldsymbol{s})-\theta_{u,d}(\boldsymbol{s},\boldsymbol{z})>\boldsymbol{0}\right\}f_{u,d}(\boldsymbol{z})\text{\rm d}z_{1}\text{\rm d}z_{2}
=\displaystyle= eu22πu22u,d(𝒛)fu,d(𝒛)dz1dz2,\displaystyle\frac{e^{-u^{2}}}{2\pi u^{2}}\iint_{\mathbb{R}^{2}}\mathcal{I}_{u,d}(\boldsymbol{z})f_{u,d}(\boldsymbol{z})\text{\rm d}z_{1}\text{\rm d}z_{2},

where u,d(𝒛):={𝒔𝑫χu,d(𝒔)θu,d(𝒔,𝒛)>𝟎}\mathcal{I}_{u,d}(\boldsymbol{z}):=\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}\chi_{u,d}(\boldsymbol{s})-\theta_{u,d}(\boldsymbol{s},\boldsymbol{z})>\boldsymbol{0}\right\} with

χu,d(𝒔)=u(ξu,d(𝒔)Ru,d(𝒔,𝟎)Ru,d1(𝟎,𝟎)ξu,d(𝟎)),𝒔𝑫,\chi_{u,d}(\boldsymbol{s})=u\left(\xi_{u,d}(\boldsymbol{s})-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\right),\quad\boldsymbol{s}\in\boldsymbol{D},
θu,d(𝒔,𝒛)=u2(𝟏Ru,d(𝒔,𝟎)Ru,d1(𝟎,𝟎)(𝟏+𝒛/u2)),𝒔𝑫,𝒛2\theta_{u,d}(\boldsymbol{s},\boldsymbol{z})=u^{2}\left(\boldsymbol{1}-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})(\boldsymbol{1}+\boldsymbol{z}/u^{2})\right),\quad\boldsymbol{s}\in\boldsymbol{D},\boldsymbol{z}\in\mathbb{R}^{2}

and

fu,d(𝒛)=11r2(d)exp(11+r(d)(u2r(d)z1z2)(z1r(d)z2)22u2(1r2(d))z222u2),𝒛2.f_{u,d}(\boldsymbol{z})=\frac{1}{\sqrt{1-r^{2}(d)}}\exp\left(\frac{1}{1+r(d)}(u^{2}r(d)-z_{1}-z_{2})-\frac{(z_{1}-r(d)z_{2})^{2}}{2u^{2}(1-r^{2}(d))}-\frac{z_{2}^{2}}{2u^{2}}\right),\quad\boldsymbol{z}\in\mathbb{R}^{2}.

Consequently, in order to show the claim it suffices to prove that for Wα(𝒔)=(2Bα(1)(s1)s1α,2Bα(2)(s2)s2α)TW_{\alpha}(\boldsymbol{s})=(\sqrt{2}B_{\alpha}^{(1)}(s_{1})-s_{1}^{\alpha},\sqrt{2}B_{\alpha}^{(2)}(s_{2})-s_{2}^{\alpha})^{T}, 𝒔𝑫\boldsymbol{s}\in\boldsymbol{D}, where Bα(1)B_{\alpha}^{(1)} and Bα(2)B_{\alpha}^{(2)} are two independent fBm’s with Hurst index α/2\alpha/2,

(6.2) limusupdA(u)|2u,d(𝒛)fu,d(𝒛)dz1dz2α2(S,0)|\displaystyle\lim_{u\to\infty}\sup_{d\geq A(u)}\left\lvert\iint_{\mathbb{R}^{2}}\mathcal{I}_{u,d}(\boldsymbol{z})f_{u,d}(\boldsymbol{z})\text{\rm d}z_{1}\text{\rm d}z_{2}-\mathcal{B}_{\alpha}^{2}(S,0)\right\rvert
=limusupdA(u)|2(u,d(𝒛)fu,d(𝒛){𝒔𝑫Wα(𝒔)+𝒛>𝟎}ez1z2)dz1dz2|\displaystyle=\lim_{u\to\infty}\sup_{d\geq A(u)}\left\lvert\iint_{\mathbb{R}^{2}}\bigg{(}\mathcal{I}_{u,d}(\boldsymbol{z})f_{u,d}(\boldsymbol{z})-\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}W_{\alpha}(\boldsymbol{s})+\boldsymbol{z}>\boldsymbol{0}\right\}e^{-z_{1}-z_{2}}\bigg{)}\text{\rm d}z_{1}\text{\rm d}z_{2}\right\rvert
=limusupdA(u)|2(u,d(𝒛)fu,d(𝒛)(𝒛)ez1z2)dz1dz2|=0,\displaystyle=\lim_{u\to\infty}\sup_{d\geq A(u)}\left\lvert\iint_{\mathbb{R}^{2}}\big{(}\mathcal{I}_{u,d}(\boldsymbol{z})f_{u,d}(\boldsymbol{z})-\mathcal{I}(\boldsymbol{z})e^{-z_{1}-z_{2}}\big{)}\text{\rm d}z_{1}\text{\rm d}z_{2}\right\rvert=0,

with

(6.3) (𝒛)\displaystyle\mathcal{I}(\boldsymbol{z}) :=\displaystyle:= {𝒔𝑫Wα(𝒔)+𝒛>𝟎}\displaystyle\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}W_{\alpha}(\boldsymbol{s})+\boldsymbol{z}>\boldsymbol{0}\right\}
(6.4) =\displaystyle= {sups[0,S]2Bα(s)sα>z1}{sups[0,S]2Bα(s)sα>z2}.\displaystyle\mathbb{P}\left\{\sup_{s_{\in}[0,S]}\sqrt{2}B_{\alpha}(s)-s^{\alpha}>-z_{1}\right\}\mathbb{P}\left\{\sup_{s\in[0,S]}\sqrt{2}B_{\alpha}(s)-s^{\alpha}>-z_{2}\right\}.

For any 𝒔,𝒕𝑫\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D},

(6.5) Cov(χu,d(𝒔),χu,d(𝒕))\displaystyle\mathrm{Cov}(\chi_{u,d}(\boldsymbol{s}),\chi_{u,d}(\boldsymbol{t}))
=u2Cov(ξu,d(𝒔)Ru,d(𝒔,𝟎)Ru,d1(𝟎,𝟎)ξu,d(𝟎),ξu,d(𝒕)Ru,d(𝒕,𝟎)Ru,d1(𝟎,𝟎)ξu,d(𝟎))\displaystyle=u^{2}\mathrm{Cov}\big{(}\xi_{u,d}(\boldsymbol{s})-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0}),\,\,\xi_{u,d}(\boldsymbol{t})-R_{u,d}(\boldsymbol{t},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\big{)}
=u2{Ru,d(𝒔,𝒕)Ru,d(𝒔,𝟎)Ru,d1(𝟎,𝟎)Ru,d(𝟎,𝒕)}\displaystyle=u^{2}\{R_{u,d}(\boldsymbol{s},\boldsymbol{t})-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})R_{u,d}(\boldsymbol{0},\boldsymbol{t})\}
=u2{(Ru,d(𝒔,𝒕)E)+(ERu,d(𝒔,𝟎))+Ru,d(𝒔,𝟎)(ERu,d1(𝟎,𝟎)Ru,d(𝟎,𝒕))}\displaystyle=u^{2}\{(R_{u,d}(\boldsymbol{s},\boldsymbol{t})-E)+(E-R_{u,d}(\boldsymbol{s},\boldsymbol{0}))+R_{u,d}(\boldsymbol{s},\boldsymbol{0})(E-R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})R_{u,d}(\boldsymbol{0},\boldsymbol{t}))\}

where EE is the 2×22\times 2 identity matrix.

Since A(u)>1A(u)>1 satisfying (6.1) tends to \infty as uu\to\infty, then by A3 we have

(6.6) limusupd>A(u),𝒔𝑫|Ru,d(𝒔,𝟎)E|=𝟎\displaystyle\lim_{u\to\infty}\sup_{d>A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert R_{u,d}(\boldsymbol{s},\boldsymbol{0})-E\right\rvert=\boldsymbol{0}

and

(6.7) limuu2supdA(u)|r(d)|limuu2logA(u)supdA(u)|r(d)|logd=0.\displaystyle\lim_{u\to\infty}u^{2}\sup_{d\geq A(u)}\left\lvert r(d)\right\rvert\leq\lim_{u\to\infty}\frac{u^{2}}{\log A(u)}\sup_{d\geq A(u)}\left\lvert r(d)\right\rvert\log d=0.

Therefore,

(6.8) limusupdA(u)u2(ERu,d1(𝟎,𝟎))=limusupdA(u)u2r(d)1r2(d)(r(d)11r(d))=𝟎.\displaystyle\lim_{u\to\infty}\sup_{d\geq A(u)}u^{2}(E-R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0}))=\lim_{u\to\infty}\sup_{d\geq A(u)}\frac{u^{2}r(d)}{1-r^{2}(d)}\begin{pmatrix}-r(d)&1\\ 1&-r(d)\end{pmatrix}=\boldsymbol{0}.

Note that by A1, (2.1) and the Uniform Convergence Theorem (see, e.g., Theorem 1.5.2 in [18]) we get

limusups[0,S]|u2(1r(s/v(u)))sα|\displaystyle\quad\lim_{u\to\infty}\sup_{s\in[0,S]}\left\lvert u^{2}(1-r(s/v(u)))-s^{\alpha}\right\rvert
limusups[0,S]|1r(s/v(u))1r(1/v(u))sα|+limusups[0,S]|1r(s/v(u))1r(1/v(u))||u2(1r(1/v(u)))1|\displaystyle\leq\lim_{u\to\infty}\sup_{s\in[0,S]}\left\lvert\frac{1-r(s/v(u))}{1-r(1/v(u))}-s^{\alpha}\right\rvert+\lim_{u\to\infty}\sup_{s\in[0,S]}\left\lvert\frac{1-r(s/v(u))}{1-r(1/v(u))}\right\rvert\left\lvert u^{2}(1-r(1/v(u)))-1\right\rvert
=0.\displaystyle=0.

Consequently,

(6.9) limusupdA(u),𝒔,𝒕𝑫|u2(ERu,d(𝒔,𝒕))(|t1s1|α00|t2s2|α)|=𝟎.\displaystyle\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}}\left\lvert u^{2}(E-R_{u,d}(\boldsymbol{s},\boldsymbol{t}))-\begin{pmatrix}\left\lvert t_{1}-s_{1}\right\rvert^{\alpha}&0\\ 0&\left\lvert t_{2}-s_{2}\right\rvert^{\alpha}\end{pmatrix}\right\rvert=\boldsymbol{0}.

Similarly,

limusupdA(u),𝒔,𝒕𝑫|u2Ru,d(𝒔,𝟎)(ERu,d1(𝟎,𝟎)Ru,d(𝟎,𝒕))(|t1|α00|t2|α)|\displaystyle\quad\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}}\left\lvert u^{2}R_{u,d}(\boldsymbol{s},\boldsymbol{0})\big{(}E-R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})R_{u,d}(\boldsymbol{0},\boldsymbol{t})\big{)}-\begin{pmatrix}\left\lvert t_{1}\right\rvert^{\alpha}&0\\ 0&\left\lvert t_{2}\right\rvert^{\alpha}\end{pmatrix}\right\rvert
limusupdA(u),𝒔,𝒕𝑫|u2(ERu,d(𝟎,𝒕))(|t1|α00|t2|α)|\displaystyle\leq\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}}\left\lvert u^{2}(E-R_{u,d}(\boldsymbol{0},\boldsymbol{t}))-\begin{pmatrix}\left\lvert t_{1}\right\rvert^{\alpha}&0\\ 0&\left\lvert t_{2}\right\rvert^{\alpha}\end{pmatrix}\right\rvert
+limusupdA(u),𝒔,𝒕𝑫|(Ru,d(𝒔,𝟎)E)[u2(ERu,d(𝟎,𝒕))]|\displaystyle\quad+\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}}\left\lvert(R_{u,d}(\boldsymbol{s},\boldsymbol{0})-E)[u^{2}(E-R_{u,d}(\boldsymbol{0},\boldsymbol{t}))]\right\rvert
+limusupdA(u),𝒔,𝒕𝑫|Ru,d(𝒔,𝟎)[u2(ERu,d1(𝟎,𝟎))]Ru,d(𝟎,𝒕)|\displaystyle\quad+\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}}\left\lvert R_{u,d}(\boldsymbol{s},\boldsymbol{0})[u^{2}(E-R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0}))]R_{u,d}(\boldsymbol{0},\boldsymbol{t})\right\rvert
(6.10) =𝟎,\displaystyle=\boldsymbol{0},

where in the last equality we have used (6.6), (6.8) and (6.9).

Substituting (6.9) and (LABEL:rcovas4) into (6.5) gives

(6.11) limusupdA(u),𝒔,𝒕𝑫|Cov(χu,d(𝒔),χu,d(𝒕))(|s1|α+|t1|α|t1s1|α00|s2|α+|t2|α|t2s2|α)|\displaystyle{\small\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}}\Bigg{|}\mathrm{Cov}(\chi_{u,d}(\boldsymbol{s}),\chi_{u,d}(\boldsymbol{t}))-\begin{pmatrix}|s_{1}|^{\alpha}+|t_{1}|^{\alpha}-\left\lvert t_{1}-s_{1}\right\rvert^{\alpha}&0\\ 0&|s_{2}|^{\alpha}+|t_{2}|^{\alpha}-\left\lvert t_{2}-s_{2}\right\rvert^{\alpha}\end{pmatrix}\Bigg{|}}
=\displaystyle= 𝟎.\displaystyle\boldsymbol{0}.

Hence, the finite-dimensional distributions of χu,d\chi_{u,d} converge to that of {2Bα(𝒔),𝒔𝑫}\{\sqrt{2}B_{\alpha}(\boldsymbol{s}),\boldsymbol{s}\in\boldsymbol{D}\} uniformly with respect to dA(u)d\geq A(u), where Bα(𝒔)=(Bα(1)(s1),Bα(2)(s2))T.B_{\alpha}(\boldsymbol{s})=(B_{\alpha}^{(1)}(s_{1}),B_{\alpha}^{(2)}(s_{2}))^{T}.

Let C(𝑫)C(\boldsymbol{D}) denote the Banach space of all continuous functions on 𝑫\boldsymbol{D} equipped with sup-norm, we now show that the measures on C(𝑫)C(\boldsymbol{D}) induced by {χu,d(𝒔),𝒔𝑫,dA(u)}\{\chi_{u,d}(\boldsymbol{s}),\boldsymbol{s}\in\boldsymbol{D},d\geq A(u)\} are uniformly tight for large uu. In fact, since 𝔼{ξu,d(𝒔)|ξu,d(𝟎)}=Ru,d(𝒔,𝟎)Ru,d1(𝟎,𝟎)ξu,d(𝟎)\mathbb{E}\left\{\xi_{u,d}(\boldsymbol{s})|\xi_{u,d}(\boldsymbol{0})\right\}=R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0}) then

𝔼{(ξu,d(𝒔)ξu,d(𝒕))T(Ru,d(𝒔,𝟎)Ru,d(𝒕,𝟎))Ru,d1(𝟎,𝟎)ξu,d(𝟎)}\displaystyle\quad\mathbb{E}\left\{(\xi_{u,d}(\boldsymbol{s})-\xi_{u,d}(\boldsymbol{t}))^{T}\left(R_{u,d}(\boldsymbol{s},\boldsymbol{0})-R_{u,d}(\boldsymbol{t},\boldsymbol{0})\right)R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\right\}
=𝔼{𝔼{(ξu,d(𝒔)ξu,d(𝒕))T|ξu,d(𝟎)}(Ru,d(𝒔,𝟎)Ru,d(𝒕,𝟎))Ru,d1(𝟎,𝟎)ξu,d(𝟎)}\displaystyle=\mathbb{E}\left\{\mathbb{E}\left\{(\xi_{u,d}(\boldsymbol{s})-\xi_{u,d}(\boldsymbol{t}))^{T}|\xi_{u,d}(\boldsymbol{0})\right\}\left(R_{u,d}(\boldsymbol{s},\boldsymbol{0})-R_{u,d}(\boldsymbol{t},\boldsymbol{0})\right)R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\right\}
=𝔼{ξu,dT(𝟎)Ru,d1(𝟎,𝟎)(Ru,d(𝟎,𝒔)Ru,d(𝟎,𝒕))(Ru,d(𝒔,𝟎)Ru,d(𝒕,𝟎))Ru,d1(𝟎,𝟎)ξu,d(𝟎)}\displaystyle=\mathbb{E}\left\{\xi_{u,d}^{T}(\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})(R_{u,d}(\boldsymbol{0},\boldsymbol{s})-R_{u,d}(\boldsymbol{0},\boldsymbol{t}))(R_{u,d}(\boldsymbol{s},\boldsymbol{0})-R_{u,d}(\boldsymbol{t},\boldsymbol{0}))R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\right\}

and thus

𝔼{χu,d(𝒔)χu,d(𝒕)2}\displaystyle\quad\mathbb{E}\left\{\lVert\chi_{u,d}(\boldsymbol{s})-\chi_{u,d}(\boldsymbol{t})\rVert^{2}\right\}
=u2[𝔼{ξu,d(𝒔)ξu,d(𝒕)2}2𝔼{(ξu,d(𝒔)ξu,d(𝒕))T(Ru,d(𝒔,𝟎)Ru,d(𝒕,𝟎))Ru,d1(𝟎,𝟎)ξu,d(𝟎)}\displaystyle=u^{2}\Big{[}\mathbb{E}\left\{\lVert\xi_{u,d}(\boldsymbol{s})-\xi_{u,d}(\boldsymbol{t})\rVert^{2}\right\}-2\mathbb{E}\left\{(\xi_{u,d}(\boldsymbol{s})-\xi_{u,d}(\boldsymbol{t}))^{T}(R_{u,d}(\boldsymbol{s},\boldsymbol{0})-R_{u,d}(\boldsymbol{t},\boldsymbol{0}))R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\right\}
+𝔼{ξu,dT(𝟎)Ru,d1(𝟎,𝟎)(Ru,d(𝟎,𝒔)Ru,d(𝟎,𝒕))(Ru,d(𝒔,𝟎)Ru,d(𝒕,𝟎))Ru,d1(𝟎,𝟎)ξu,d(𝟎)}]\displaystyle\quad+\mathbb{E}\left\{\xi_{u,d}^{T}(\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})(R_{u,d}(\boldsymbol{0},\boldsymbol{s})-R_{u,d}(\boldsymbol{0},\boldsymbol{t}))(R_{u,d}(\boldsymbol{s},\boldsymbol{0})-R_{u,d}(\boldsymbol{t},\boldsymbol{0}))R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\right\}\Big{]}
=u2[𝔼{ξu,d(𝒔)ξu,d(𝒕)2}𝔼{(Ru,d(𝒔,𝟎)Ru,d(𝒕,𝟎))Ru,d1(𝟎,𝟎)ξu,d(𝟎)2}]\displaystyle=u^{2}\Big{[}\mathbb{E}\left\{\lVert\xi_{u,d}(\boldsymbol{s})-\xi_{u,d}(\boldsymbol{t})\rVert^{2}\right\}-\mathbb{E}\left\{\lVert(R_{u,d}(\boldsymbol{s},\boldsymbol{0})-R_{u,d}(\boldsymbol{t},\boldsymbol{0}))R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\xi_{u,d}(\boldsymbol{0})\rVert^{2}\right\}\Big{]}
u2Eξu,d(𝒔)ξu,d(𝒕)2\displaystyle\leq u^{2}E{\lVert\xi_{u,d}(\boldsymbol{s})-\xi_{u,d}(\boldsymbol{t})\rVert^{2}}
=2u2[1r(|t1s1|/v(u))+1r(|t2s2|/v(u))].\displaystyle=2u^{2}[1-r(\left\lvert t_{1}-s_{1}\right\rvert/v(u))+1-r(\left\lvert t_{2}-s_{2}\right\rvert/v(u))].

Moreover, by (2.1) and Potter’s Theorem (see, e.g., [18][Theorem 1.5.6]), for large enough uu there exists some constant C>1C>1 such that

u2(1r(s/v(u)))=u2(1r(1/v(u)))1r(s/v(u))1r(1/v(u))C|s|α/2\displaystyle u^{2}(1-r(s/v(u)))=u^{2}(1-r(1/v(u)))\frac{1-r(s/v(u))}{1-r(1/v(u))}\leq C\left\lvert s\right\rvert^{\alpha/2}

holds for all s[0,S]s\in[0,S]. Hence, for large enough uu we get

(6.12) supdA(u)𝔼{χu,d(𝒔)χu,d(𝒕)2}2C(|t1s1|α/2+|t2s2|α/2)\displaystyle\sup_{d\geq A(u)}\mathbb{E}\left\{\lVert\chi_{u,d}(\boldsymbol{s})-\chi_{u,d}(\boldsymbol{t})\rVert^{2}\right\}\leq 2C(|t_{1}-s_{1}|^{\alpha/2}+|t_{2}-s_{2}|^{\alpha/2})

for any 𝒔,𝒕𝑫\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}, implying the uniform tightness of the measures induced by {χu,d(𝒔),𝒔𝑫,dA(u)}\{\chi_{u,d}(\boldsymbol{s}),\boldsymbol{s}\in\boldsymbol{D},d\geq A(u)\}. This together with (6.11) implies that χu,d\chi_{u,d} converges weakly, as uu\to\infty, to 2Bα(𝒔),𝒔𝑫\sqrt{2}B_{\alpha}(\boldsymbol{s}),\boldsymbol{s}\in\boldsymbol{D} uniformly for dA(u)d\geq A(u).

Further, by (6.8) and (6.9) we have

limusupdA(u),𝒔𝑫|u2(ERu,d(𝒔,𝟎)Ru,d1(𝟎,𝟎))(|s1|α00|s2|α)|\displaystyle\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert u^{2}\big{(}E-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\big{)}-\begin{pmatrix}\left\lvert s_{1}\right\rvert^{\alpha}&0\\ 0&\left\lvert s_{2}\right\rvert^{\alpha}\end{pmatrix}\right\rvert
limusupdA(u),𝒔𝑫|u2(ERu,d(𝒔,𝟎))(|s1|α00|s2|α)|\displaystyle\leq\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert u^{2}(E-R_{u,d}(\boldsymbol{s},\boldsymbol{0}))-\begin{pmatrix}\left\lvert s_{1}\right\rvert^{\alpha}&0\\ 0&\left\lvert s_{2}\right\rvert^{\alpha}\end{pmatrix}\right\rvert
+limusupdA(u),𝒔𝑫|Ru,d(𝒔,𝟎)[u2(ERu,d1(𝟎,𝟎))]|\displaystyle\quad+\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert R_{u,d}(\boldsymbol{s},\boldsymbol{0})[u^{2}(E-R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0}))]\right\rvert
=𝟎\displaystyle=\boldsymbol{0}

and thus for any 𝒛2\boldsymbol{z}\in\mathbb{R}^{2}

limusupdA(u),𝒔𝑫|θu,d(𝒔)(|s1|αz1,|s2|αz2)T|\displaystyle\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert\theta_{u,d}(\boldsymbol{s})-(|s_{1}|^{\alpha}-z_{1},|s_{2}|^{\alpha}-z_{2})^{T}\right\rvert
limusupdA(u),𝒔𝑫|[u2(ERu,d(𝒔,𝟎)Ru,d1(𝟎,𝟎))(|s1|α00|s2|α)][11]|\displaystyle\leq\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert\Bigg{[}u^{2}\big{(}E-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\big{)}-\begin{pmatrix}\left\lvert s_{1}\right\rvert^{\alpha}&0\\ 0&\left\lvert s_{2}\right\rvert^{\alpha}\end{pmatrix}\Bigg{]}\begin{bmatrix}1\\ 1\end{bmatrix}\right\rvert
+limusupdA(u),𝒔𝑫|(ERu,d(𝒔,𝟎)Ru,d1(𝟎,𝟎))𝒛|\displaystyle\quad+\quad\lim_{u\to\infty}\sup_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert\big{(}E-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\big{)}\boldsymbol{z}\right\rvert
=𝟎.\displaystyle=\boldsymbol{0}.

Therefore, for each 𝒛2\boldsymbol{z}\in\mathbb{R}^{2}, the probability measures on C(𝑫)C(\boldsymbol{D}) induced by {χu,d(𝒔)θu,d(𝒔,𝒛),𝒔𝑫}\{\chi_{u,d}(\boldsymbol{s})-\theta_{u,d}(\boldsymbol{s},\boldsymbol{z}),\boldsymbol{s}\in\boldsymbol{D}\} converge weakly, as uu\to\infty, to that induced by {Wα(𝒔)+𝒛,𝒔𝑫}\{W_{\alpha}(\boldsymbol{s})+\boldsymbol{z},\boldsymbol{s}\in\boldsymbol{D}\} uniformly with respect to dA(u)d\geq A(u). Then, by the continuous mapping theorem, (6.4) and the fact that the set of discontinuity points of cumulative distribution function of sups[0,S]2Bα(s)sα\sup_{s_{\in}[0,S]}\sqrt{2}B_{\alpha}(s)-s^{\alpha} consists of at most of one point (see, e.g., Theorem 7.1 in [19] or related Lemma 4.4 in [20]), we get

limusupdA(u)|u,d(𝒛)(𝒛)|=0\displaystyle\lim_{u\to\infty}\sup_{d\geq A(u)}\left\lvert\mathcal{I}_{u,d}(\boldsymbol{z})-\mathcal{I}(\boldsymbol{z})\right\rvert=0

for almost all 𝒛2\boldsymbol{z}\in\mathbb{R}^{2}, where (𝒛)\mathcal{I}(\boldsymbol{z}) is defined in (6.3). Further, by (6.7) we know

limusupdA(u)|fu,d(𝒛)ez1z2|=0,𝒛2,\lim_{u\to\infty}\sup_{d\geq A(u)}\left\lvert f_{u,d}(\boldsymbol{z})-e^{-z_{1}-z_{2}}\right\rvert=0,\quad\forall\,\boldsymbol{z}\in\mathbb{R}^{2},

and thus for almost all 𝒛2\boldsymbol{z}\in\mathbb{R}^{2}

(6.14) limusupdA(u)|u,d(𝒛)fu,d(𝒛)(𝒛)ez1z2|=0.\displaystyle\lim_{u\to\infty}\sup_{d\geq A(u)}\left\lvert\mathcal{I}_{u,d}(\boldsymbol{z})f_{u,d}(\boldsymbol{z})-\mathcal{I}(\boldsymbol{z})e^{-z_{1}-z_{2}}\right\rvert=0.

Therefore, to verify (6.2), we have to put the limit into integral. In the following, we look for an integrable upper bound for supdA(u)u,d(𝒛)fu,d(𝒛)\sup_{d\geq A(u)}\mathcal{I}_{u,d}(\boldsymbol{z})f_{u,d}(\boldsymbol{z}). We first give a lower bound for infdA(u),𝒔𝑫θu,d(𝒔,𝒛)\inf_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\theta_{u,d}(\boldsymbol{s},\boldsymbol{z}). Let ε(<1/2)\varepsilon(<1/2) be a positive constant. In view of (6), we know that, for sufficiently large uu

supdA(u),𝒔𝑫|ERu,d(𝒔,𝟎)Ru,d1(𝟎,𝟎)|(εεεε),\displaystyle\sup_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\left\lvert E-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0})\right\rvert\leq\begin{pmatrix}\varepsilon&\varepsilon\\ \varepsilon&\varepsilon\end{pmatrix},

and thus

infdA(u),𝒔𝑫θu,d(𝒔,𝒛)\displaystyle\inf_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\theta_{u,d}(\boldsymbol{s},\boldsymbol{z}) =\displaystyle= infdA(u),𝒔𝑫{u2(ERu,d(𝒔,𝟎)Ru,d1(𝟎,𝟎))𝟏+(ERu,d(𝒔,𝟎)Ru,d1(𝟎,𝟎))𝒛𝒛}\displaystyle\inf_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\{u^{2}(E-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0}))\boldsymbol{1}+(E-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0}))\boldsymbol{z}-\boldsymbol{z}\}
\displaystyle\geq 𝟏𝒛+infdA(u),𝒔𝑫{(ERu,d(𝒔,𝟎)Ru,d1(𝟎,𝟎))𝒛}\displaystyle-\boldsymbol{1}-\boldsymbol{z}+\inf_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\{(E-R_{u,d}(\boldsymbol{s},\boldsymbol{0})R_{u,d}^{-1}(\boldsymbol{0},\boldsymbol{0}))\boldsymbol{z}\}
\displaystyle\geq 𝟏𝒛(εεεε)|𝒛|:=h(𝒛),𝒛2.\displaystyle-\boldsymbol{1}-\boldsymbol{z}-\begin{pmatrix}\varepsilon&\varepsilon\\ \varepsilon&\varepsilon\end{pmatrix}\left\lvert\boldsymbol{z}\right\rvert:=h(\boldsymbol{z}),\quad\boldsymbol{z}\in\mathbb{R}^{2}.

Let {𝒆k,k=1,2,3}\{\boldsymbol{e}_{k},k=1,2,3\} denotes (1,1)T(1,1)^{T}, (0,1)T(0,1)^{T} and (1,0)T(1,0)^{T}, respectively. By Cauchy-Schwartz inequality and (6.12), for large enough uu

(6.15) supdA(u)𝔼{(𝒆kT(χu,d(𝒔)χu,d(𝒕)))2}\displaystyle\sup_{d\geq A(u)}\mathbb{E}\left\{\left(\boldsymbol{e}_{k}^{T}(\chi_{u,d}(\boldsymbol{s})-\chi_{u,d}(\boldsymbol{t}))\right)^{2}\right\} \displaystyle\leq supdA(u)2𝔼{χu,d(𝒔)χu,d(𝒕)2}\displaystyle\sup_{d\geq A(u)}2\mathbb{E}\left\{\lVert\chi_{u,d}(\boldsymbol{s})-\chi_{u,d}(\boldsymbol{t})\rVert^{2}\right\}
\displaystyle\leq 4C(|t1s1|α/2+|t2s2|α/2),k=1,2,3\displaystyle 4C(|t_{1}-s_{1}|^{\alpha/2}+|t_{2}-s_{2}|^{\alpha/2}),\quad k=1,2,3

holds for any 𝒔,𝒕𝑫\boldsymbol{s},\boldsymbol{t}\in\boldsymbol{D}. Thus, by Sudakov-Fernique inequality (see, e.g., [21][Theorem 2.9]), we have

(6.16) supdA(u)𝔼{sup𝒔𝑫𝒆kTχu,d(𝒔)}𝔼{sup𝒔𝑫i=122CBα/2(i)(si)}:=C1<,k=1,2,3,\displaystyle\sup_{d\geq A(u)}\mathbb{E}\left\{\sup_{\boldsymbol{s}\in\boldsymbol{D}}\boldsymbol{e}_{k}^{T}\chi_{u,d}(\boldsymbol{s})\right\}\leq\mathbb{E}\left\{\sup_{\boldsymbol{s}\in\boldsymbol{D}}\sum_{i=1}^{2}2\sqrt{C}B_{\alpha/2}^{(i)}(s_{i})\right\}:=C_{1}<\infty,\quad k=1,2,3,

where Bα/2(i)B_{\alpha/2}^{(i)}’s are independent fBm’s with Hurst index α/4\alpha/4. Then, for all large enough uu,

supdA(u)u,d(𝒛)\displaystyle\sup_{d\geq A(u)}\mathcal{I}_{u,d}(\boldsymbol{z}) =\displaystyle= supdA(u){𝒔𝑫χu,d(𝒔)θu,d(𝒔,𝒛)>𝟎}\displaystyle\sup_{d\geq A(u)}\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}\chi_{u,d}(\boldsymbol{s})-\theta_{u,d}(\boldsymbol{s},\boldsymbol{z})>\boldsymbol{0}\right\}
\displaystyle\leq supdA(u){𝒔𝑫χu,d(𝒔)>infdA(u),𝒔𝑫θu,d(𝒔,𝒛)}\displaystyle\sup_{d\geq A(u)}\mathbb{P}\left\{\exists_{\boldsymbol{s}\in\boldsymbol{D}}\chi_{u,d}(\boldsymbol{s})>\inf_{d\geq A(u),\boldsymbol{s}\in\boldsymbol{D}}\theta_{u,d}(\boldsymbol{s},\boldsymbol{z})\right\}
\displaystyle\leq supdA(u){sup𝒔𝑫𝒆kTχu,d(𝒔)>𝒆kTh(𝒛)}\displaystyle\sup_{d\geq A(u)}\mathbb{P}\left\{\sup_{\boldsymbol{s}\in\boldsymbol{D}}\boldsymbol{e}_{k}^{T}\chi_{u,d}(\boldsymbol{s})>\boldsymbol{e}_{k}^{T}h(\boldsymbol{z})\right\}
\displaystyle\leq supdA(u)exp((𝒆kTh(𝒛)𝔼{sup𝒔𝑫𝒆kTχu,d(𝒔)})22Var𝒔𝑫𝒆kTχu,d(𝒔))\displaystyle\sup_{d\geq A(u)}\exp\left(-\frac{\left(\boldsymbol{e}_{k}^{T}h(\boldsymbol{z})-\mathbb{E}\left\{\sup_{\boldsymbol{s}\in\boldsymbol{D}}\boldsymbol{e}_{k}^{T}\chi_{u,d}(\boldsymbol{s})\right\}\right)^{2}}{2\text{Var}_{\boldsymbol{s}\in\boldsymbol{D}}\boldsymbol{e}_{k}^{T}\chi_{u,d}(\boldsymbol{s})}\right)
\displaystyle\leq exp(C2(𝒆kTh(𝒛)C1)2),𝒛𝒁k,k=1,2,3,\displaystyle\exp\left(-C_{2}\left(\boldsymbol{e}_{k}^{T}h(\boldsymbol{z})-C_{1}\right)^{2}\right),\quad\boldsymbol{z}\in\boldsymbol{Z}_{k},k=1,2,3,

where (6) follows from Borell-TIS inequality (see, e.g., [22][Theorem 2.1.1]), the last inequality follows by (6.15)-(6.16) with C2=(16CSα/2)1C_{2}=(16CS^{\alpha/2})^{-1}, and

𝒁1={(z1,z2)|z1<0,z2<0,(2ε1)(z1+z2)>2+C1},\boldsymbol{Z}_{1}=\{(z_{1},z_{2})|z_{1}<0,z_{2}<0,(2\varepsilon-1)(z_{1}+z_{2})>2+C_{1}\},
𝒁2={(z1,z2)|z1>0,z2<0,(ε1)z2εz1>1+C1},\boldsymbol{Z}_{2}=\{(z_{1},z_{2})|z_{1}>0,z_{2}<0,(\varepsilon-1)z_{2}-\varepsilon z_{1}>1+C_{1}\},
𝒁3={(z1,z2)|z1<0,z2>0,(ε1)z1εz2>1+C1}.\boldsymbol{Z}_{3}=\{(z_{1},z_{2})|z_{1}<0,z_{2}>0,(\varepsilon-1)z_{1}-\varepsilon z_{2}>1+C_{1}\}.

Therefore,

supdA(u)u,d(𝒛)g(𝒛):={exp(C2(𝒆kTh(𝒛)C1)2),𝒛𝒁k,k=1,2,3,1,𝒛2\k=13𝒁k,\displaystyle\sup_{d\geq A(u)}\mathcal{I}_{u,d}(\boldsymbol{z})\leq g(\boldsymbol{z}):=\left\{\begin{array}[]{ll}\exp\left(-C_{2}\left(\boldsymbol{e}_{k}^{T}h(\boldsymbol{z})-C_{1}\right)^{2}\right),&\boldsymbol{z}\in\boldsymbol{Z}_{k},k=1,2,3,\\ 1,&\quad\boldsymbol{z}\in\mathbb{R}^{2}\backslash\bigcup_{k=1}^{3}\boldsymbol{Z}_{k},\\ \end{array}\right.

holds for sufficiently large uu. Moreover, by (6.7)

supdA(u)fu,d(𝒛)ez1+z2\displaystyle\quad\sup_{d\geq A(u)}f_{u,d}(\boldsymbol{z})e^{z_{1}+z_{2}}
=supdA(u)11r2(d)exp(u2r(d)1+r(d)+2u2r(d)(1r(d))(z1+z2)(z122r(d)z1z2+z22)2u2(1r2(d)))\displaystyle=\sup_{d\geq A(u)}\frac{1}{\sqrt{1-r^{2}(d)}}\exp\left(\frac{u^{2}r(d)}{1+r(d)}+\frac{2u^{2}r(d)(1-r(d))(z_{1}+z_{2})-(z_{1}^{2}-2r(d)z_{1}z_{2}+z_{2}^{2})}{2u^{2}(1-r^{2}(d))}\right)
32supdA(u)exp(u2r(d)1+r(d)+1+r(d)2(z2z1)21r(d)2(z1+z22u2r(d))2+2u4r2(d)(1r(d))2u2(1r2(d)))\displaystyle\leq\frac{3}{2}\sup_{d\geq A(u)}\exp\left(\frac{u^{2}r(d)}{1+r(d)}+\frac{-\frac{1+r(d)}{2}(z_{2}-z_{1})^{2}-\frac{1-r(d)}{2}(z_{1}+z_{2}-2u^{2}r(d))^{2}+2u^{4}r^{2}(d)(1-r(d))}{2u^{2}(1-r^{2}(d))}\right)
32supdA(u)eu2r(d)2,𝒛2\displaystyle\leq\frac{3}{2}\sup_{d\geq A(u)}e^{u^{2}r(d)}\leq 2,\quad\boldsymbol{z}\in\mathbb{R}^{2}

holds for all large enough uu, and thus

supdA(u)u,d(𝒛)fu,d(𝒛)2g(𝒛)ez1z2,𝒛2.\displaystyle\sup_{d\geq A(u)}\mathcal{I}_{u,d}(\boldsymbol{z})f_{u,d}(\boldsymbol{z})\leq 2g(\boldsymbol{z})e^{-z_{1}-z_{2}},\quad\boldsymbol{z}\in\mathbb{R}^{2}.

We now show that g(𝒛)ez1z2g(\boldsymbol{z})e^{-z_{1}-z_{2}} is integrable on 2\mathbb{R}^{2}. In fact,

2g(𝒛)ez1z2dz1dz2=(𝒁1+𝒁2+𝒁3+2\k=13𝒁k)g(𝒛)ez1z2dz1dz2,\displaystyle\iint_{\mathbb{R}^{2}}g(\boldsymbol{z})e^{-z_{1}-z_{2}}\text{\rm d}z_{1}\text{\rm d}z_{2}=\left(\iint_{\boldsymbol{Z}_{1}}+\iint_{\boldsymbol{Z}_{2}}+\iint_{\boldsymbol{Z}_{3}}+\iint_{\mathbb{R}^{2}\backslash\bigcup_{k=1}^{3}\boldsymbol{Z}_{k}}\right)g(\boldsymbol{z})e^{-z_{1}-z_{2}}\text{\rm d}z_{1}\text{\rm d}z_{2},

where

𝒁1g(𝒛)ez1z2dz1dz2\displaystyle\iint_{\boldsymbol{Z}_{1}}g(\boldsymbol{z})e^{-z_{1}-z_{2}}\text{\rm d}z_{1}\text{\rm d}z_{2} \displaystyle\leq 00exp(C2((2ε1)z1+(2ε1)z22C1)2z1z2)dz1dz2\displaystyle\int_{-\infty}^{0}\int_{-\infty}^{0}\exp\left(-C_{2}\left((2\varepsilon-1)z_{1}+(2\varepsilon-1)z_{2}-2-C_{1}\right)^{2}-z_{1}-z_{2}\right)\text{\rm d}z_{1}\text{\rm d}z_{2}
\displaystyle\leq (0exp(C2(2ε1)2z12+(2C2(2ε1)(2+C1)1)z1)dz1)2\displaystyle\left(\int_{-\infty}^{0}\exp\left(-C_{2}(2\varepsilon-1)^{2}z_{1}^{2}+\left(2C_{2}(2\varepsilon-1)(2+C_{1})-1\right)z_{1}\right)\text{\rm d}z_{1}\right)^{2}
<\displaystyle< ,\displaystyle\infty,
𝒁2g(𝒛)ez1z2dz1dz2\displaystyle\iint_{\boldsymbol{Z}_{2}}g(\boldsymbol{z})e^{-z_{1}-z_{2}}\text{\rm d}z_{1}\text{\rm d}z_{2} =\displaystyle= 𝒁3g(𝒛)ez1z2dz1dz2\displaystyle\iint_{\boldsymbol{Z}_{3}}g(\boldsymbol{z})e^{-z_{1}-z_{2}}\text{\rm d}z_{1}\text{\rm d}z_{2}
=\displaystyle= 0(εz1+1+C1(ε1)exp(C2((ε1)z2εz11C1)2z2)dz2)ez1dz1\displaystyle\int^{\infty}_{0}\left(\int_{-\infty}^{\frac{\varepsilon z_{1}+1+C_{1}}{(\varepsilon-1)}}\exp\left(-C_{2}\left((\varepsilon-1)z_{2}-\varepsilon z_{1}-1-C_{1}\right)^{2}-z_{2}\right)\text{\rm d}z_{2}\right)e^{-z_{1}}\text{\rm d}z_{1}
=\displaystyle= e1+C11ε1ε0e(ε1ε1)z1dz10exp(C2z22z2ε1)dz2<,\displaystyle\frac{e^{\frac{1+C_{1}}{1-\varepsilon}}}{1-\varepsilon}\int^{\infty}_{0}e^{(\frac{\varepsilon}{1-\varepsilon}-1)z_{1}}\text{\rm d}z_{1}\int_{0}^{\infty}\exp\left(-C_{2}z_{2}^{2}-\frac{z_{2}}{\varepsilon-1}\right)\text{\rm d}z_{2}<\infty,

since ε<1/2\varepsilon<1/2, and

2\k=13𝒁kg(𝒛)ez1z2dz1dz2\displaystyle\iint_{\mathbb{R}^{2}\backslash\bigcup_{k=1}^{3}\boldsymbol{Z}_{k}}g(\boldsymbol{z})e^{-z_{1}-z_{2}}\text{\rm d}z_{1}\text{\rm d}z_{2} \displaystyle\leq (z1<0,z2<0,z1+z22+C12ε1+20εz1+C1+1ε1)ez1z2dz1dz2\displaystyle\left(\iint_{z_{1}<0,z_{2}<0,z_{1}+z_{2}\geq\frac{2+C_{1}}{2\varepsilon-1}}+2\int_{0}^{\infty}\int^{\infty}_{\frac{\varepsilon z_{1}+C_{1}+1}{\varepsilon-1}}\right)e^{-z_{1}-z_{2}}\text{\rm d}z_{1}\text{\rm d}z_{2}
\displaystyle\leq (2+C112ε)2e2+C112ε+2e1+C11ε0e(ε1ε1)z1dz1<.\displaystyle\left(\frac{2+C_{1}}{1-2\varepsilon}\right)^{2}e^{\frac{2+C_{1}}{1-2\varepsilon}}+2e^{\frac{1+C_{1}}{1-\varepsilon}}\int^{\infty}_{0}e^{(\frac{\varepsilon}{1-\varepsilon}-1)z_{1}}\text{\rm d}z_{1}<\infty.

Consequently, (6.2) follows by the dominated convergence theorem and (6.14). This completes the proof. \Box

Lemma 6.2.

Let X(t),t0X(t),t\geq 0 be a centered stationary Gaussian process with unit variance and covariance function satisfying A1 and A3. Let v(u)v(u), α(x)\mathcal{B}_{\alpha}(x) and m(u)m(u) be defined in (2.1), (2.3) and (2.5) respectively. Then for A(u)A(u) such that

(6.19) limuA(u)v(u)=andlimuA(u)m(u)=0\displaystyle\lim_{u\to\infty}A(u)v(u)=\infty\quad\textrm{and}\quad\lim_{u\to\infty}\frac{A(u)}{m(u)}=0

and any x0x\geq 0 we have

(6.20) {Lu[0,A(u)]>x}α(x)A(u)v(u)Ψ(u),u.\displaystyle\mathbb{P}\left\{L_{u}^{\ast}[0,A(u)]>x\right\}\sim\mathcal{B}_{\alpha}(x)A(u)v(u)\Psi(u),\quad u\to\infty.

Proof of Lemma 6.2 We follow the argument used in the proof of Theorem 2.1 in [6]. Let A(u)A(u) satisfy (6.19), for any S>1S>1 define

Δk=[kS/v(u),(k+1)S/v(u)],k=0,,Nu\Delta_{k}=[kS/v(u),(k+1)S/v(u)],\quad k=0,\ldots,N_{u}

with Nu=A(u)v(u)/SN_{u}=\lfloor A(u)v(u)/S\rfloor, i.e., the integer part of A(u)v(u)/SA(u)v(u)/S. By stationarity of XX, we have for all uu positive and x0x\geq 0

I1(u){Lu[0,T]>x}I2(u),\displaystyle I_{1}(u)\leq\mathbb{P}\left\{L_{u}^{*}[0,T]>x\right\}\leq I_{2}(u),

where

I1(u)=(Nu1){LuΔ0>x}0i<kNu1qi,k(u),\displaystyle I_{1}(u)=(N_{u}-1)\mathbb{P}\left\{L_{u}^{*}\Delta_{0}>x\right\}-\sum_{0\leq i<k\leq N_{u}-1}q_{i,k}(u),
I2(u)=(Nu+1){LuΔ0>x}+0i<kNuqi,k(u),\displaystyle I_{2}(u)=(N_{u}+1)\mathbb{P}\left\{L_{u}^{*}\Delta_{0}>x\right\}+\sum_{0\leq i<k\leq N_{u}}q_{i,k}(u),

with qi,k(u)={suptΔiX(t)>u,suptΔkX(t)>u}.q_{i,k}(u)=\mathbb{P}\left\{\sup_{t\in\Delta_{i}}X(t)>u,\sup_{t\in\Delta_{k}}X(t)>u\right\}. By Theorem 5.1 in [6] and (2.3), we have

(6.21) limSlimuNu{LuΔ0>x}A(u)v(u)Ψ(u)=α(x)\displaystyle\lim_{S\to\infty}\lim_{u\to\infty}\frac{N_{u}\mathbb{P}\left\{L_{u}^{*}\Delta_{0}>x\right\}}{A(u)v(u)\Psi(u)}=\mathcal{B}_{\alpha}(x)

for any x0x\geq 0. Therefore, it suffices to show that the double sum is negligible with respect to A(u)v(u)Ψ(u)A(u)v(u)\Psi(u) as uu\to\infty and then as SS\to\infty.

Let ε(<2)\varepsilon^{\ast}(<2) be the positive root of equation x2(2α)x32α=0x^{2}-(2-\alpha)x-\frac{3}{2}\alpha=0 and put β=inft1{1r(t)},\beta=\inf_{t\geq 1}\{1-r(t)\}, which by A3 is positive. Define

A0(u)=0,A1(u)=uε2αA(u),A2(u)=1A(u),A3(u)=eβu2/8A(u),A4(u)=A(u)\displaystyle A_{0}(u)=0,\,A_{1}(u)=u^{\frac{\varepsilon^{*}-2}{\alpha}}\wedge A(u),\,A_{2}(u)=1\wedge A(u),\,A_{3}(u)=e^{\beta u^{2}/8}\wedge A(u),\,A_{4}(u)=A(u)

and

Λl(u)={(i,k):1i+1<kNu,Al1(u)<(ki1)S/v(u)Al(u),l=1,2,3,4}.\Lambda_{l}(u)=\{(i,k):1\leq i+1<k\leq N_{u},A_{l-1}(u)<(k-i-1)S/v(u)\leq A_{l}(u),\,l=1,2,3,4\}.

Then

0i<kNuqi,k(u)\displaystyle\sum_{0\leq i<k\leq N_{u}}q_{i,k}(u) =\displaystyle= 0i<Nuqi,i+1(u)+l=14(i,k)Λl(u)qi,k(u)\displaystyle\sum_{0\leq i<N_{u}}q_{i,i+1}(u)+\sum_{l=1}^{4}\sum_{(i,k)\in\Lambda_{l}(u)}q_{i,k}(u)
:=\displaystyle:= Q0(u)+l=14Ql(u).\displaystyle Q_{0}(u)+\sum_{l=1}^{4}Q_{l}(u).

According to (4.7)-(4.9) in [6] we know that

(6.23) lim supuQ2(u)A(u)v(u)Ψ(u)=0,\displaystyle\limsup_{u\to\infty}\frac{Q_{2}(u)}{A(u)v(u)\Psi(u)}=0,
(6.24) limSlim supuQ1(u)A(u)v(u)Ψ(u)=0,\displaystyle\lim_{S\to\infty}\limsup_{u\to\infty}\frac{Q_{1}(u)}{A(u)v(u)\Psi(u)}=0,

and

(6.25) limSlim supuQ0(u)A(u)v(u)Ψ(u)=0.\displaystyle\lim_{S\to\infty}\limsup_{u\to\infty}\frac{Q_{0}(u)}{A(u)v(u)\Psi(u)}=0.

Next, by stationarity of XX, for sufficiently large uu

sup(i,k)Λ3(u)𝔼{supsΔi,tΔk(X(s)+X(t))}2𝔼{sups[0,1]X(s)}=:C3<,\displaystyle\sup_{(i,k)\in\Lambda_{3}(u)}\mathbb{E}\left\{\sup_{s\in\Delta_{i},t\in\Delta_{k}}(X(s)+X(t))\right\}\leq 2\mathbb{E}\left\{\sup_{s\in[0,1]}X(s)\right\}=:C_{3}<\infty,
sup(i,k)Λ3(u),sΔi,tΔkVar(X(s)+X(t))\displaystyle\sup_{(i,k)\in\Lambda_{3}(u),s\in\Delta_{i},t\in\Delta_{k}}\text{Var}(X(s)+X(t)) =\displaystyle= 42inf(i,k)Λ3(u),(s,t)Δi×Δk{1r(ts)}\displaystyle 4-2\inf_{(i,k)\in\Lambda_{3}(u),(s,t)\in\Delta_{i}\times\Delta_{k}}\{1-r(t-s)\}
\displaystyle\leq 42β.\displaystyle 4-2\beta.

Then, by Borell-TIS inequality we have for large enough uu

sup(i,k)Λ3(u)qi,k(u)\displaystyle\sup_{(i,k)\in\Lambda_{3}(u)}q_{i,k}(u) \displaystyle\leq sup(i,k)Λ3(u){supsΔi,tΔkX(s)+X(t)>2u}\displaystyle\sup_{(i,k)\in\Lambda_{3}(u)}\mathbb{P}\left\{\sup_{s\in\Delta_{i},t\in\Delta_{k}}X(s)+X(t)>2u\right\}
\displaystyle\leq exp((2uC3)22(42β))\displaystyle\exp\left(-\frac{(2u-C_{3})^{2}}{2(4-2\beta)}\right)
\displaystyle\leq exp(1+β/22(uC3/2)2),\displaystyle\exp\left(-\frac{1+\beta/2}{2}(u-C_{3}/2)^{2}\right),

and thus

(6.26) lim supuQ3(u)A(u)v(u)Ψ(u)\displaystyle\limsup_{u\to\infty}\frac{Q_{3}(u)}{A(u)v(u)\Psi(u)} \displaystyle\leq lim supuNuA3(u)v(u)SA(u)v(u)Ψ(u)exp(1+β/22(uC3/2)2)\displaystyle\limsup_{u\to\infty}\frac{N_{u}A_{3}(u)v(u)}{SA(u)v(u)\Psi(u)}\exp\left(-\frac{1+\beta/2}{2}(u-C_{3}/2)^{2}\right)
\displaystyle\leq lim supu2πuv(u)S2exp(β8u2+C3u(1+β/2))=0.\displaystyle\limsup_{u\to\infty}\frac{\sqrt{2\pi}uv(u)}{S^{2}}\exp\left(-\frac{\beta}{8}u^{2}+C_{3}u(1+\beta/2)\right)=0.

Further, since eβu2/8e^{\beta u^{2}/8} satisfies (6.1), then by Lemma 6.1 and stationarity of XX,

Q4(u)2Nu2Ψ2(u)α2(S,0)Q_{4}(u)\leq 2N^{2}_{u}\Psi^{2}(u)\mathcal{B}_{\alpha}^{2}(S,0)

holds for uu sufficiently large. Therefore,

lim supuQ4(u)A(u)v(u)Ψ(u)\displaystyle\limsup_{u\to\infty}\frac{Q_{4}(u)}{A(u)v(u)\Psi(u)} \displaystyle\leq lim supu2Nu2Ψ2(u)α2(S,0)A(u)v(u)Ψ(u)\displaystyle\limsup_{u\to\infty}\frac{2N_{u}^{2}\Psi^{2}(u)\mathcal{B}_{\alpha}^{2}(S,0)}{A(u)v(u)\Psi(u)}
\displaystyle\leq lim supu2α2(S,0)S2α(0)A(u)m(u)=0,\displaystyle\limsup_{u\to\infty}\frac{2\mathcal{B}_{\alpha}^{2}(S,0)}{S^{2}\mathcal{B}_{\alpha}(0)}\frac{A(u)}{m(u)}=0,

where the last equality follows by (6.19).

Consequently, substituting (6.23)-(6) into (6) yields

limSlim supu1A(u)v(u)Ψ(u)0i<kNuqi,k(u)=0,\displaystyle\lim_{S\to\infty}\limsup_{u\to\infty}\frac{1}{A(u)v(u)\Psi(u)}\sum_{0\leq i<k\leq N_{u}}q_{i,k}(u)=0,

which together with (6.21) completes the proof. \Box

Corollary 6.1.

If XX, v(u)v(u), α(x)\mathcal{B}_{\alpha}(x), m(u)m(u) and A(u)A(u) are given as in Lemma 6.2, then for any x0x\geq 0 and ε(0,1)\varepsilon\in(0,1) there exists δ>0\delta>0 such that

(6.27) lim infuinft[A(u),δm(u)]{Lu[0,t]>x}tα(x)v(u)Ψ(u)1ε\displaystyle\liminf_{u\to\infty}\inf_{t\in[A(u),\delta m(u)]}\frac{\mathbb{P}\left\{L_{u}^{*}[0,t]>x\right\}}{t\mathcal{B}_{\alpha}(x)v(u)\Psi(u)}\geq 1-\varepsilon

and

(6.28) lim supusupt[A(u),δm(u)]{Lu[0,t]>x}tα(x)v(u)Ψ(u)1+ε.\displaystyle\limsup_{u\to\infty}\sup_{t\in[A(u),\delta m(u)]}\frac{\mathbb{P}\left\{L_{u}^{*}[0,t]>x\right\}}{t\mathcal{B}_{\alpha}(x)v(u)\Psi(u)}\leq 1+\varepsilon.

Proof of Corollary 6.1 Let x0x\geq 0 be fixed, recalling (5.2) we have that, for arbitrary ε>0\varepsilon>0, there exists some δ>0\delta>0 such that

(6.29) (1ε/4){Y(t)>x}tα¯(x)(1+ε/4),t(0,δ).\displaystyle(1-\varepsilon/4)\leq\frac{\mathbb{P}\left\{Y(t)>x\right\}}{t\overline{\mathcal{F}_{\alpha}}(x)}\leq(1+\varepsilon/4),\quad t\in(0,\delta).

For such ε\varepsilon and δ\delta, suppose that (6.27) does not hold. Then, there exist two sequences {un,n}\{u_{n},n\in\mathbb{N}\} and {tn,n}\{t_{n},n\in\mathbb{N}\} such that unu_{n}\to\infty as nn\to\infty and

(6.30) {Lun[0,tn]>x}tnα(x)v(un)Ψ(un)<1ε,tn[A(un),δm(un)],n.\displaystyle\frac{\mathbb{P}\left\{L_{u_{n}}^{*}[0,t_{n}]>x\right\}}{t_{n}\mathcal{B}_{\alpha}(x)v(u_{n})\Psi(u_{n})}<1-\varepsilon,\quad t_{n}\in[A(u_{n}),\delta m(u_{n})],n\in\mathbb{N}.

Putting t^n=tn/m(un){\hat{t}}_{n}=t_{n}/m(u_{n}), by (2.5) and (3.4), we get

(6.30) {Lun[0,t^nm(un)]>x}t^nα¯(x)<1ε,t^n[A(un)/m(un),δ],n.\frac{\mathbb{P}\left\{L_{u_{n}}^{*}[0,{\hat{t}}_{n}m(u_{n})]>x\right\}}{{\hat{t}}_{n}\overline{\mathcal{F}_{\alpha}}(x)}<1-\varepsilon,\quad{\hat{t}}_{n}\in[A(u_{n})/m(u_{n}),\delta],n\in\mathbb{N}.

Since sequence {t^n,n}\{{\hat{t}}_{n},n\in\mathbb{N}\} is bounded, then there exists a convergent subsequence {t^nk,k}\{{\hat{t}}_{n_{k}},k\in\mathbb{N}\} such that limkt^nk0\lim_{k\to\infty}{\hat{t}}_{n_{k}}\geq 0. If limkt^nk>0\lim_{k\to\infty}{\hat{t}}_{n_{k}}>0, then by Corollary 5.1

{Lunk[0,t^nkm(unk)]>x}{Y(t^nk)>x}>1ε/4\displaystyle\frac{\mathbb{P}\left\{L_{u_{n_{k}}}^{*}[0,{\hat{t}}_{n_{k}}m(u_{n_{k}})]>x\right\}}{\mathbb{P}\left\{Y({\hat{t}}_{n_{k}})>x\right\}}>1-\varepsilon/4

holds for sufficiently large kk, which together with (6.29) implies

{Lunk[0,t^nkm(unk)]>x}t^nkα¯(x)>(1ε/4)2.\displaystyle\frac{\mathbb{P}\left\{L_{u_{n_{k}}}^{*}[0,{\hat{t}}_{n_{k}}m(u_{n_{k}})]>x\right\}}{{\hat{t}}_{n_{k}}\overline{\mathcal{F}_{\alpha}}(x)}>(1-\varepsilon/4)^{2}.

This however contradicts (6.30). If limkt^nk=0\lim_{k\to\infty}{\hat{t}}_{n_{k}}=0, then

limktnkv(unk)limkA(unk)v(unk)=andlimktnkm(unk)=limkt^nk=0,\displaystyle\lim_{k\to\infty}t_{n_{k}}v(u_{n_{k}})\geq\lim_{k\to\infty}A(u_{n_{k}})v(u_{n_{k}})=\infty\quad\textrm{and}\quad\lim_{k\to\infty}\frac{t_{n_{k}}}{m(u_{n_{k}})}=\lim_{k\to\infty}{\hat{t}}_{n_{k}}=0,

and thus by Lemma 6.2

{Lunk[0,tnk]>x}tnkα(x)v(unk)Ψ(unk)>1ε/4\displaystyle\frac{\mathbb{P}\left\{L_{u_{n_{k}}}^{*}[0,t_{n_{k}}]>x\right\}}{t_{n_{k}}\mathcal{B}_{\alpha}(x)v(u_{n_{k}})\Psi(u_{n_{k}})}>1-\varepsilon/4

holds for sufficiently large kk. This contradicts (6.30). An analogous argument can be used to verify (6.28). This completes the proof. \Box

Acknowledgement: The authors would like to thank Enkelejd Hashorva for his numerous valuable remarks on all the steps of preparation of the manuscript. K.D. was partially supported by NCN Grant No 2018/31/B/ST1/00370 (2019-2022). X.P. thanks National Natural Science Foundation of China (11701070,71871046) for partial financial support. Financial support from the Swiss National Science Foundation Grant 200021-175752/1 is also kindly acknowledged.

References

  • [1] H. Guérin and J.-F. Renaud, “On the distribution of cumulative Parisian ruin,” Insurance Math. Econom., vol. 73, pp. 116–123, 2017.
  • [2] D. Landriault, B. Li, and A. L. Mohamed, “On occupation times in the red of Lévy risk models,” http://arXiv:1903.03721, 2019.
  • [3] S. M. Berman, “Sojourns above a high level for a Gaussian process with a point of maximum variance,” Comm. Pure Appl. Math., vol. 38, no. 5, pp. 519–528, 1985.
  • [4] S. M. Berman, “Extreme sojourns of a Gaussian process with a point of maximum variance,” Probability Theory and Related Fields, vol. 74, no. 1, pp. 113–124, 1987.
  • [5] S. M. Berman, Sojourns and extremes of stochastic processes. The Wadsworth & Brooks/Cole Statistics/Probability Series, Pacific Grove, CA: Wadsworth & Brooks/Cole Advanced Books & Software, 1992.
  • [6] K. Dȩbicki, Z. Michna, and X. Peng, “Approximation of sojourn times of Gaussian processes,” Methodology and Computing in Applied Probability, vol. 21, no. 4, pp. 1183–1213, 2019.
  • [7] K. Dȩbicki, P. Liu, and Z. Michna, “Sojourn times of Gaussian processes with trend,” Accepted for publication in Journal of Theoretical Probability, 2019.
  • [8] S. Fotopoulos and Y. Luo, “Subordinated Gaussian processes, the log-return principles,” Washington State University, Washington, 2011.
  • [9] H. Geman, D. B. Madan, and M. Yor, “Time changes for Lévy processes,” Mathematical Finance, vol. 11, no. 1, pp. 79–96, 2001.
  • [10] F. den Hollander, S. N. Majumdar, J. M. Meylahn, and H. Touchette, “Properties of additive functionals of Brownian motion with resetting,” Journal of Physics A: Mathematical and Theoretical, vol. 52, p. 175001, Apr 2019.
  • [11] B. Zwart, S. Borst, and K. Dȩbicki, “Subexponential asymptotics of hybrid fluid and ruin models,” The Annals of Applied Probability, vol. 15, no. 1A, pp. 500–517, 2005.
  • [12] J. Hüsler, V. Piterbarg, and E. Rumyantseva, “Extremes of Gaussian processes with a smooth random variance,” Stochastic Processes and their Applications, vol. 121, no. 11, pp. 2592–2605, 2011.
  • [13] J. Hüsler, V. Piterbarg, and Y. Zhang, “Extremes of Gaussian processes with random variance,” Electronic Journal of Probability, vol. 16, pp. 1254–1280, 2011.
  • [14] Z. Tan and E. Hashorva, “Exact tail asymptotics of the supremum of strongly dependent Gaussian processes over a random interval,” Lithuanian Mathematical Journal, vol. 53, no. 1, pp. 91–102, 2013.
  • [15] K. Dȩbicki, J. Farkas, and E. Hashorva, “Extremes of randomly scaled Gumbel risks,” Journal of Mathematical Analysis and Applications, vol. 458, no. 1, pp. 30–42, 2018.
  • [16] M. Arendarczyk and K. Dȩbicki, “Exact asymptotics of supremum of a stationary Gaussian process over a random interval,” Statistics and Probability Letters, vol. 82, no. 3, pp. 645–652, 2012.
  • [17] V. I. Piterbarg, Asymptotic methods in the theory of Gaussian processes and fields, vol. 148 of Translations of Mathematical Monographs. Providence, RI: American Mathematical Society, 1996.
  • [18] N. H. Bingham, C. M. Goldie, and J. L. Teugels, Regular variation, vol. 27 of Encycolpedia of Mathematics and its Applications. Cambridge: Cambridge University Press, 1989.
  • [19] J.-M. Azaïs and M. Wschebor, Level sets and extrema of random processes and fields. John Wiley & Sons, 2009.
  • [20] K. Dȩbicki, E. Hashorva, and L. Wang, “Extremes of vector-valued Gaussian processes,” arXiv e-prints: arXiv:1911.06350, 2019.
  • [21] R. J. Adler, An introduction to continuity, extrema, and related topics for general Gaussian processes, vol. 12 of Lecture Notes-Monograph Series. Institute of Mathematical Statistics, Hayward, California, 1990.
  • [22] R. J. Adler and J. E. Taylor, Random fields and geometry. Springer Monographs in Mathematics, New York: Springer, 2007.