1 Introduction
The Hermite processes of order $k=1,2,\dots $ are defined as multiple Wiener–Itô integrals
where $d=\frac{1}{2}-\frac{1-H}{k}\in (\frac{1}{2}-\frac{1}{2k},\frac{1}{2})$ and $\frac{1}{2}<H<1$ (the prime ${^{\prime }}$ on the integral sign shows that one does not integrate on the diagonals $x_{i}=x_{j}$, $i\ne j$). They are self-similar processes with stationary increments (see [8, 26]).
(1)
\[{Z_{H}^{k}}(t):={\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\int _{0}^{t}}\Bigg(\prod \limits_{i=1}^{k}{(s-y_{i})_{+}^{d-1}}\Bigg)\hspace{0.1667em}ds\hspace{0.1667em}B(dy_{1})\dots B(dy_{k}),\]In this paper, we introduce a new class of stochastic processes, which we call tempered Hermite processes. Tempered Hermite processes modify the kernel of ${Z_{H}^{k}}$ by multiplying an exponential tempering factor $\lambda >0$ such that they are well defined for Hurst parameter $H>\frac{1}{2}$. Tempered Hermite processes are not self-similar processes, but they have a scaling property, involving both the time scale and the tempering parameter. The scaling property enable us to show that the tempered Hermite processes are the weak convergence limits of certain discrete chaos processes.
2 Tempered Hermite process
Let $B=\{B(t),t\in \mathbb{R}\}$ be a real-valued Brownian motion on the real line, a process with stationary independent increments such that $B(t)$ has a Gaussian distribution with mean zero and variance $|t|$ for all $t\in \mathbb{R}$. Then the Wiener–Itô integrals
\[I_{k}(f):={\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}f(x_{1},\dots ,x_{k})B(dx_{1})\dots B(dx_{k})\]
are defined for all functions $f\in {L}^{2}({\mathbb{R}}^{k})$. The prime ′ on the integral sign shows that one does not integrate on the diagonals $x_{i}=x_{j}$, $i\ne j$. See, for example, [12, Chapter 4].Definition 1.
Let $H>\frac{1}{2}$ and $\lambda >0$. The process
where $(x)_{+}=xI(x>0)$ and $d=\frac{1}{2}-\frac{1-H}{k}\in (\frac{1}{2}-\frac{1}{2k},\infty )$, is called a tempered Hermite process of order k.
(2)
\[{Z_{H,\lambda }^{k}}(t):={\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\int _{0}^{t}}\prod \limits_{i=1}^{k}\big({(s-y_{i})_{+}^{d-1}}{e}^{-\lambda (s-y_{i})_{+}}\big)\hspace{0.1667em}ds\hspace{0.1667em}B(dy_{1})\dots B(dy_{k}),\]The next lemma shows that ${Z_{H,\lambda }^{k}}(t)$, given by (2), is well defined for any $t\ge 0$.
Proof.
To show that $h_{t}(y_{1},\dots ,y_{k})$ is in ${L}^{2}({\mathbb{R}}^{k})$, we write
where we applied the standard integral formula [13, p. 344]
for $|\arg \beta |<\pi $, Re $\mu >0$, Re $\nu >0$. Here $K_{\nu }(x)$ is the modified Bessel function of the second kind (see [1, Chapter 9]). Next, we need to show that
which is finite, and, consequently, from (4) and (6) we get
where we applied the fact that $K_{\nu }(z)=K_{-\nu }(z)$ and $k_{\nu }(z)<{z}^{-\nu }{2}^{\nu -1}\varGamma (\nu )$ for $z>0$. Hence, from (4) and (7) it follows that
Since $K_{0}(z)\sim -\log (z)$ as $z\to 0$ (see [1, Eq. 9.6.8], we have
Now, we find an upper bound for $I_{2}$. It can be shown that $\frac{K_{\nu }(x)}{K_{\nu }(y)}>{e}^{y-x}$ for $0<x<y$ and an arbitrary real number ν (see [4]). Therefore,
From (8), (9), and (10) we can see that
for $\epsilon >0$, and this shows that
(4)
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \int _{{\mathbb{R}}^{k}}h_{t}{(y_{1},\dots ,y_{k})}^{2}\hspace{0.1667em}dy_{1}\dots dy_{k}\\{} & \displaystyle \hspace{1em}=\int _{{\mathbb{R}}^{k}}\Bigg[{\int _{0}^{t}}{\int _{0}^{t}}\prod \limits_{i=1}^{k}{(s_{1}-y_{i})_{+}^{d-1}}{e}^{-\lambda (s_{1}-y_{i})_{+}}{(s_{2}-y_{i})_{+}^{d-1}}\\{} & \displaystyle \hspace{2em}\times {e}^{-\lambda (s_{2}-y_{i})_{+}}\hspace{0.1667em}ds_{1}\hspace{0.1667em}ds_{2}\Bigg]\hspace{0.1667em}dy_{1}\dots dy_{k}\\{} & \displaystyle \hspace{1em}=2{\int _{0}^{t}}ds_{1}{\int _{s_{1}}^{t}}ds_{2}\Bigg[\int _{{\mathbb{R}}^{k}}\prod \limits_{i=1}^{k}{(s_{1}-y_{i})_{+}^{d-1}}{e}^{-\lambda (s_{1}-y_{i})_{+}}{(s_{2}-y_{i})_{+}^{d-1}}\\{} & \displaystyle \hspace{2em}\times {e}^{-\lambda (s_{2}-y_{i})_{+}}\hspace{0.1667em}dy_{1}\dots dy_{k}\Bigg]\\{} & \displaystyle \hspace{1em}=2{\int _{0}^{t}}ds{\int _{0}^{t-s}}du\hspace{0.1667em}\Bigg[\int _{{\mathbb{R}_{+}^{k}}}\prod \limits_{i=1}^{k}{w_{i}^{d-1}}{e}^{-\lambda w_{i}}{(w_{i}+u)}^{d-1}{e}^{-\lambda (w_{i}+u)}\hspace{0.1667em}dw_{1}\dots dw_{k}\Bigg]\\{} & \displaystyle \hspace{2em}(s=s_{1},u=s_{2}-s_{1},w_{i}=s_{1}-y_{i})\\{} & \displaystyle \hspace{1em}=2{\int _{0}^{t}}ds{\int _{0}^{t-s}}{e}^{-\lambda uk}\hspace{0.1667em}du\hspace{0.1667em}{\bigg[\int _{\mathbb{R}_{+}}{w}^{d-1}{(w+u)}^{d-1}{e}^{-2\lambda w}\hspace{0.1667em}dw\bigg]}^{k}\\{} & \displaystyle \hspace{1em}=2{\int _{0}^{t}}ds{\int _{0}^{t-s}}{e}^{-\lambda uk}{u}^{k(2d-1)}\hspace{0.1667em}du\hspace{0.1667em}{\bigg[\int _{\mathbb{R}_{+}}{x}^{d-1}{(x+1)}^{d-1}{e}^{-2\lambda ux}\hspace{0.1667em}dx\bigg]}^{k}\\{} & \displaystyle \hspace{1em}=2{\int _{0}^{t}}ds{\int _{0}^{t-s}}{e}^{-\lambda uk}{u}^{k(2d-1)}\hspace{0.1667em}du{\bigg[\frac{\varGamma (d)}{\sqrt{\pi }}{\bigg(\frac{1}{2\lambda u}\bigg)}^{d-\frac{1}{2}}{e}^{\lambda u}K_{\frac{1}{2}-d}(\lambda u)\bigg]}^{k}\\{} & \displaystyle \hspace{1em}=2{\bigg[\frac{\varGamma (d)}{\sqrt{\pi }{(2\lambda )}^{d-\frac{1}{2}}}\bigg]}^{k}{\int _{0}^{t}}ds{\int _{0}^{t-s}}{\big[{u}^{d-\frac{1}{2}}K_{\frac{1}{2}-d}(\lambda u)\big]}^{k}\hspace{0.1667em}du\\{} & \displaystyle \hspace{1em}=2{\bigg[\frac{\varGamma (d)}{\sqrt{\pi }{2}^{d-\frac{1}{2}}{\lambda }^{2d-1}}\bigg]}^{k}{\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{\big[{z}^{d-\frac{1}{2}}K_{\frac{1}{2}-d}(z)\big]}^{k}\hspace{0.1667em}dz,\end{array}\](5)
\[{\int _{0}^{\infty }}{x}^{\nu -1}{(x+\beta )}^{\nu -1}{e}^{-\mu x}\hspace{0.1667em}dx=\frac{1}{\sqrt{\pi }}{\bigg(\frac{\beta }{\mu }\bigg)}^{\nu -\frac{1}{2}}{e}^{\frac{\beta \mu }{2}}\varGamma (\nu )K_{\frac{1}{2}-\nu }\bigg(\frac{\beta \mu }{2}\bigg)\]
\[{\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{\big[{z}^{d-\frac{1}{2}}K_{\frac{1}{2}-d}(z)\big]}^{k}\hspace{0.1667em}dz\]
is finite for $d>\frac{1}{2}-\frac{1}{2k}$ (equivalently, for $H>\frac{1}{2}$). First, suppose $\frac{1}{2}-\frac{1}{2k}<d<\frac{1}{2}$ (or $\frac{1}{2}<H<1$). Since $k_{\nu }(z)<{z}^{-\nu }{2}^{\nu -1}\varGamma (\nu )$ for $z>0$ (Theorem 3.1 in [11]), we have
(6)
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle {\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{\big[{z}^{d-\frac{1}{2}}K_{\frac{1}{2}-d}(z)\big]}^{k}\hspace{0.1667em}dz\\{} & \displaystyle \hspace{1em}\le {\bigg[{2}^{-(\frac{1}{2}+d)}\varGamma \bigg(\frac{1}{2}-d\bigg)\bigg]}^{k}{\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{z}^{k(2d-1)}\hspace{0.1667em}dz\\{} & \displaystyle \hspace{1em}=\frac{{[{\lambda }^{2d-1}{2}^{-(\frac{1}{2}+d)}\varGamma (\frac{1}{2}-d)]}^{k}\lambda }{(k(2d-1)+1)(k(2d-1)+2)}{t}^{2kd-k+2},\end{array}\]
\[\int _{{\mathbb{R}}^{k}}h_{t}{(y_{1},\dots ,y_{k})}^{2}\hspace{0.1667em}dy_{1}\dots dy_{k}<\frac{2\lambda {[\frac{\varGamma (d)\varGamma (\frac{1}{2}-d)}{\sqrt{\pi }{2}^{2d}}]}^{k}}{(k(2d-1)+1)(k(2d-1)+2)}{t}^{2kd-k+2}\]
for $\frac{1}{2}-\frac{1}{2k}<d<\frac{1}{2}$. Next, suppose $d>\frac{1}{2}$ (equivalently $H>1$). In this case,
(7)
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle {\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{\big[{z}^{d-\frac{1}{2}}K_{\frac{1}{2}-d}(z)\big]}^{k}\hspace{0.1667em}dz={\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{\big[{z}^{d-\frac{1}{2}}K_{d-\frac{1}{2}}(z)\big]}^{k}\hspace{0.1667em}dz\\{} & \displaystyle \hspace{1em}\le {\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{\bigg[{2}^{d-\frac{3}{2}}\varGamma \bigg(d-\frac{1}{2}\bigg)\bigg]}^{k}\hspace{0.1667em}dz\le \frac{\lambda {[{2}^{d-\frac{3}{2}}\varGamma (d-\frac{1}{2})]}^{k}}{2}{t}^{2},\end{array}\]
\[\int _{{\mathbb{R}}^{k}}h_{t}{(y_{1},\dots ,y_{k})}^{2}\hspace{0.1667em}dy_{1}\dots dy_{k}<{\bigg[\frac{\varGamma (d)\varGamma (d-\frac{1}{2})}{2\sqrt{\pi }{\lambda }^{2d-1}}\bigg]}^{k}\lambda \hspace{0.1667em}{t}^{2}\]
for $d>\frac{1}{2}$. Finally, let $d=\frac{1}{2}$ (equivalently, $H=1$). Consider
(8)
\[\begin{array}{r@{\hskip0pt}l}\displaystyle {\int _{0}^{\lambda (t-s)}}{\big(K_{0}(z)\big)}^{k}\hspace{0.1667em}dz& \displaystyle ={\int _{0}^{\eta }}{\big(K_{0}(z)\big)}^{k}\hspace{0.1667em}dz+{\int _{\eta }^{\lambda (t-s)}}{\big(K_{0}(z)\big)}^{k}\hspace{0.1667em}dz\\{} & \displaystyle :=I_{1}+I_{2}.\end{array}\](9)
\[\begin{array}{r@{\hskip0pt}l}\displaystyle I_{1}& \displaystyle ={\int _{0}^{\eta }}{\bigg(\frac{K_{0}(z)}{\log (z)}\log (z)\bigg)}^{k}\hspace{0.1667em}dz\le {(1+\epsilon )}^{k}{\int _{0}^{\eta }}{\big(-\log (z)\big)}^{k}\hspace{0.1667em}dz\\{} & \displaystyle ={(1+\epsilon )}^{k}{\int _{-\log (\eta )}^{+\infty }}{w}^{k}{e}^{-w}\hspace{0.1667em}dz\le {(1+\epsilon )}^{k}{\int _{0}^{+\infty }}{w}^{k}{e}^{-w}\hspace{0.1667em}dz\\{} & \displaystyle ={(1+\epsilon )}^{k}\varGamma (k+1).\end{array}\](10)
\[\begin{array}{r@{\hskip0pt}l}\displaystyle I_{2}& \displaystyle ={\int _{\eta }^{\lambda (t-s)}}{\big(K_{0}(z)\big)}^{k}\hspace{0.1667em}dz<{\int _{\eta }^{\lambda (t-s)}}{\big(K_{0}(\eta ){e}^{\eta -z}\big)}^{k}\hspace{0.1667em}dz\hspace{0.1667em}{\big[K_{0}(\eta ){e}^{\eta }\big]}^{k}\big(\lambda (t-s)-\eta \big).\end{array}\]
\[{\int _{0}^{\lambda (t-s)}}{\big(K_{0}(z)\big)}^{k}\hspace{0.1667em}dz<{(1+\epsilon )}^{k}\varGamma (k+1)+{\big[K_{0}(\eta ){e}^{\eta }\big]}^{k}\big(\lambda (t-s)-\eta \big)\]
and hence
(11)
\[{\int _{0}^{t}}ds{\int _{0}^{\lambda (t-s)}}{\big[K_{\frac{1}{2}-d}(z)\big]}^{k}\hspace{0.1667em}dz<\big({(1+\epsilon )}^{k}\varGamma (k+1)\big)t\hspace{0.1667em}+\hspace{0.1667em}{\big[K_{0}(\eta ){e}^{\eta }\big]}^{k}\bigg(\frac{\lambda {t}^{2}}{2}-\eta t\bigg)\]
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \int _{{\mathbb{R}}^{k}}h_{t}{(y_{1},\dots ,y_{k})}^{2}\hspace{0.1667em}dy_{1}\dots dy_{k}\\{} & \displaystyle \hspace{1em}<2{\bigg[\frac{\varGamma (d)}{\sqrt{\pi }{\lambda }^{2d-1}}{2}^{d-\frac{1}{2}}\bigg]}^{k}\bigg[\big({(1+\epsilon )}^{k}\varGamma (k+1)\big)t+{\big[K_{0}(\eta ){e}^{\eta }\big]}^{k}\bigg(\frac{\lambda {t}^{2}}{2}-\eta t\bigg)\bigg]\end{array}\]
for $d=\frac{1}{2}$ ($H=1$), which completes the proof. □The next result shows that although a tempered Hermite process is not a self-similar process, it does have a nice scaling property. Here the symbol ≜ indicates the equivalence of finite-dimensional distributions.
Proposition 1.
The process ${Z_{H,\lambda }^{k}}$ given by (2) has stationary increments such that
for any scale factor $c>0$.
(12)
\[\big\{{Z_{H,\lambda }^{k}}(ct)\big\}_{t\in \mathbb{R}}\triangleq \big\{{c}^{H}{Z_{H,c\lambda }^{k}}(t)\big\}_{t\in \mathbb{R}}\]Proof.
Since $B(dy)$ has the control measure $m(dy)={\sigma }^{2}\hspace{0.1667em}dy$, the random measure $B(c\hspace{0.1667em}dy)$ has the control measure ${c}^{1/2}{\sigma }^{2}\hspace{0.1667em}dy$. Given $t_{j}$, $j=1,\dots ,n$, by the change of variables $s=c{s^{\prime }}$ and $y_{i}=c{y^{\prime }_{i}}$ for $i=1\dots ,k$ we have
\[\begin{array}{r@{\hskip0pt}l}\displaystyle {Z_{H,\lambda }^{k}}(ct_{j})& \displaystyle \hspace{0.1667em}=\hspace{0.1667em}\int _{{\mathbb{R}}^{k}}\hspace{-0.1667em}{\int _{0}^{ct_{j}}}\Bigg(\prod \limits_{i=1}^{k}{(s\hspace{0.1667em}-\hspace{0.1667em}y_{i})_{+}^{-(\frac{1}{2}+\frac{1-H}{k})}}{e}^{-\lambda (s-y_{i})_{+}}\Bigg)\hspace{0.1667em}ds\hspace{0.1667em}B(dy_{1})\dots B(dy_{k})\\{} & \displaystyle \hspace{0.1667em}=\hspace{0.1667em}\int _{{\mathbb{R}}^{k}}\hspace{-0.1667em}{\int _{0}^{t_{j}}}\Bigg(\prod \limits_{i=1}^{k}{\big(c{s^{\prime }}-\hspace{0.1667em}c{y_{i}^{{^{\prime }}}}\big)_{+}^{-(\frac{1}{2}+\frac{1-H}{k})}}{e}^{-\lambda (c{s^{\prime }}-c{y_{i}^{{^{\prime }}}})_{+}}\hspace{-0.1667em}\Bigg)c\hspace{0.1667em}d{s^{\prime }}\hspace{0.1667em}B\big(cd{y^{\prime }_{1}}\big)\dots B\big(cd{y^{\prime }_{k}}\big)\\{} & \displaystyle \hspace{0.1667em}\triangleq \hspace{0.1667em}{c}^{H}\int _{\mathbb{R}}\hspace{-0.1667em}{\int _{0}^{t_{j}}}\big({\big({s^{\prime }}\hspace{0.1667em}-\hspace{0.1667em}{y^{\prime }}\big)_{+}^{-(\frac{1}{2}+\frac{1-H}{k})}}{e}^{-\lambda c({s^{\prime }}-{y^{\prime }})_{+}}\big)\hspace{0.1667em}d{s^{\prime }}\hspace{0.1667em}B\big(d{y^{\prime }_{1}}\big)\dots B\big(d{y^{\prime }_{k}}\big)\\{} & \displaystyle \hspace{0.1667em}=\hspace{0.1667em}{c}^{H}{Z_{H,c\lambda }^{k}}(t_{j}),\end{array}\]
so that (12) holds. Suppose now that $s_{j}<t_{j}$ and change the variables $x={x^{\prime }}+s_{j}$, $y_{i}=s_{j}+{y^{\prime }_{i}}$ (for $j=1,\dots ,n$) to get
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \big({Z_{H,\lambda }^{k}}(t_{j})-{Z_{H,\lambda }^{k}}(s_{j})\big)\\{} & \displaystyle \hspace{1em}=\int _{{\mathbb{R}}^{k}}{\int _{s_{j}}^{t_{j}}}\Bigg(\prod \limits_{i=1}^{k}{(x-y_{i})_{+}^{-(\frac{1}{2}+\frac{1-H}{k})}}{e}^{-\lambda (x-y_{i})_{+}}\Bigg)\hspace{0.1667em}dx\hspace{0.1667em}B(dy_{1})\dots B(dy_{k})\\{} & \displaystyle \hspace{1em}=\int _{{\mathbb{R}}^{k}}{\int _{0}^{t_{j}-s_{j}}}\Bigg(\prod \limits_{i=1}^{k}{\big({x^{\prime }}+s_{j}-y_{i}\big)_{+}^{-(\frac{1}{2}+\frac{1-H}{k})}}{e}^{-\lambda ({x^{\prime }}+s_{j}-y_{i})_{+}}\Bigg)\hspace{0.1667em}d{x^{\prime }}\hspace{0.1667em}B(dy_{1})\dots B(dy_{k})\\{} & \displaystyle \hspace{1em}\triangleq \int _{{\mathbb{R}}^{k}}{\int _{0}^{t_{j}-s_{j}}}\Bigg(\prod \limits_{i=1}^{k}{\big({x^{\prime }_{i}}-{y^{\prime }_{i}}\big)_{+}^{-(\frac{1}{2}+\frac{1-H}{k})}}{e}^{-\lambda ({x^{\prime }_{i}}-{y^{\prime }_{i}})_{+}}\Bigg)\hspace{0.1667em}d{x^{\prime }}\hspace{0.1667em}B\big(d{y^{\prime }_{1}}\big)\dots B\big(d{y^{\prime }_{k}}\big)\\{} & \displaystyle \hspace{1em}={Z_{H,\lambda }^{k}}(t_{j}-s_{j}),\end{array}\]
which shows that a tempered Hermite process of order k has stationary increments. □Proof.
According to the proof of Lemma 1,
where $c_{1}$ and $c_{2}$ are some constants. Kolmogorov’s continuity criterion states that a stochastic process $X(t)$ has a continuous version if there exist some positive constants p, β, and c such that
Apply (14) for tempered Hermite process ${Z_{H,\lambda }^{k}}(t)$ by taking $p=1$, $\beta =\min \{1,2H-1\}$, and $c=\min \{c_{1},c_{2}\}$ to get the desired result. □
(13)
\[\mathbb{E}{\big|{Z_{H,\lambda }^{k}}(t)-{Z_{H,\lambda }^{k}}(s)\big|}^{2}\le \left\{\begin{array}{l@{\hskip10.0pt}l}c_{1}|t-s{|}^{2H}\hspace{1em}& \frac{1}{2}<H<1,\\{} c_{2}|t-s{|}^{2}\hspace{1em}& H>1,\end{array}\right.\]We next compute the covariance function of ${Z_{H,\lambda }^{k}}(t)$. Unlike Hermite processes, the covariance function of a tempered Hermite process is different for different $k\ge 1$.
Proposition 3.
The process ${Z_{H,\lambda }^{k}}$ given by (2) has the covariance function
\[R(t,s)=2{\bigg[\frac{\varGamma (d)}{\sqrt{\pi }{(2\lambda )}^{d-\frac{1}{2}}}\bigg]}^{k}{\int _{0}^{t}}{\int _{0}^{s}}{\big[|u-v{|}^{d-\frac{1}{2}}K_{\frac{1}{2}-d}\big(\lambda |u-v|\big)\big]}^{k}\hspace{0.1667em}dv\hspace{0.1667em}du\]
for $\lambda >0$ and $d>\frac{1}{2}-\frac{1}{2k}$ (equivalently, $H>\frac{1}{2}$).
Proof.
By applying the Fubini theorem and the isometry of multiple Wiener–Itô integrals we have
\[\begin{array}{r@{\hskip0pt}l}\displaystyle R(t,s)& \displaystyle =2\int _{{\mathbb{R}}^{k}}\Bigg({\int _{0}^{t}}{\int _{0}^{s}}\prod \limits_{i=1}^{k}{(u-y_{i})_{+}^{d-1}}{(v-y_{i})_{+}^{d-1}}\\{} & \displaystyle \hspace{1em}\times {e}^{-\lambda (u-y_{i})_{+}}{e}^{-\lambda (v-y_{i})_{+}}\hspace{0.1667em}dv\hspace{0.1667em}du\Bigg)\hspace{0.1667em}dy_{1}\dots dy_{k}\\{} & \displaystyle =2{\int _{0}^{t}}{\int _{0}^{s}}\int _{{\mathbb{R}}^{k}}\Bigg[\prod \limits_{i=1}^{k}{(u-y_{i})_{+}^{d-1}}{(v-y_{i})_{+}^{d-1}}\\{} & \displaystyle \hspace{1em}\times {e}^{-\lambda (u-y_{i})_{+}}{e}^{-\lambda (v-y_{i})_{+}}\hspace{0.1667em}dy_{1}\dots dy_{k}\Bigg]\hspace{0.1667em}dv\hspace{0.1667em}du\\{} & \displaystyle =2{\int _{0}^{t}}{\int _{0}^{s}}{\bigg[\int _{\mathbb{R}}{(u-y)_{+}^{d-1}}{(v-y)_{+}^{d-1}}{e}^{-\lambda (u-y)_{+}}{e}^{-\lambda (v-y)_{+}}\hspace{0.1667em}dy\bigg]}^{k}\hspace{0.1667em}dv\hspace{0.1667em}du\\{} & \displaystyle =2{\int _{0}^{t}}{\int _{0}^{s}}{\Bigg[{\int _{-\infty }^{\min (u,v)}}{(u-y)}^{d-1}{(v-y)}^{d-1}{e}^{-\lambda (u-y)}{e}^{-\lambda (v-y)}\hspace{0.1667em}dy\Bigg]}^{k}\hspace{0.1667em}dv\hspace{0.1667em}du\\{} & \displaystyle =2{\int _{0}^{t}}{\int _{0}^{s}}{\Bigg[{\int _{0}^{+\infty }}{w}^{d-1}{\big(|u-v|+w\big)}^{d-1}{e}^{-\lambda w}{e}^{-\lambda (|u-v|+w)}\hspace{0.1667em}dw\Bigg]}^{k}\hspace{0.1667em}dv\hspace{0.1667em}du\\{} & \displaystyle =2{\int _{0}^{t}}{\int _{0}^{s}}{e}^{-\lambda k|u-v|}\hspace{0.1667em}|u-v{|}^{k(2d-1)}\\{} & \displaystyle \hspace{1em}\times {\Bigg[{\int _{0}^{+\infty }}{x}^{-(\frac{1}{2}+\frac{1-H}{k})}{(x+1)}^{-(\frac{1}{2}+\frac{1-H}{k})}{e}^{-2\lambda |u-v|x}\hspace{0.1667em}dx\Bigg]}^{k}\hspace{0.1667em}dv\hspace{0.1667em}du\\{} & \displaystyle =2{\int _{0}^{t}}{\int _{0}^{s}}{e}^{-\lambda k|u-v|}\hspace{0.1667em}|u-v{|}^{k(2d-1)}\\{} & \displaystyle \hspace{1em}\times {\bigg[\frac{\varGamma (d)}{\sqrt{\pi }}{\bigg(\frac{1}{2\lambda |u-v|}\bigg)}^{d-\frac{1}{2}}{e}^{\lambda |u-v|}K_{\frac{1}{2}-d}\big(\lambda |u-v|\big)\bigg]}^{k}\hspace{0.1667em}dv\hspace{0.1667em}du\\{} & \displaystyle =2{\bigg[\frac{\varGamma (d)}{\sqrt{\pi }{(2\lambda )}^{d-\frac{1}{2}}}\bigg]}^{k}{\int _{0}^{t}}{\int _{0}^{s}}{\big[|u-v{|}^{d-\frac{1}{2}}K_{\frac{1}{2}-d}\big(\lambda |u-v|\big)\big]}^{k}\hspace{0.1667em}dv\hspace{0.1667em}du\end{array}\]
for any $H>\frac{1}{2}$ and $\lambda >0$, and hence we get the desired result. □Let $\widehat{B}_{1}$ and $\widehat{B}_{2}$ be independent Gaussian random measures with $\widehat{B}_{1}(A)=\widehat{B}_{1}(-A)$, $\widehat{B}_{2}(A)=-\widehat{B}_{2}(-A)$, and $\mathbb{E}[{(\widehat{B}_{i}(A))}^{2}]=m(A)/2$, where $m(dx)={\sigma }^{2}\hspace{0.1667em}dx$, and define the complex-valued Gaussian random measure $\widehat{B}=\widehat{B}_{1}+i\widehat{B}_{2}$.
Proposition 4.
Let $H>\frac{1}{2}$ and $\lambda >0$. The process ${Z_{H,\lambda }^{k}}$ given by (2) has the spectral domain representation
where $\widehat{B}(\cdot )$ is a complex-valued Gaussian random measure, and $C_{H,k}={(\frac{\varGamma (\frac{1}{2}-\frac{1-H}{k})}{\sqrt{2\pi }})}^{k}$ is a constant depending on H and k. The double prime ${^{\prime\prime }}$ on the integral indicates that one does not integrate on the diagonals $\omega _{i}=\omega _{j}$, $i\ne j$.
(15)
\[\begin{array}{r@{\hskip0pt}l}\displaystyle {Z_{H,\lambda }^{k}}(t)& \displaystyle =C_{H,k}{\int _{{\mathbb{R}}^{k}}^{{^{\prime\prime }}}}\frac{{e}^{it(\omega _{1}+\cdots +\omega _{k})}-1}{i(\omega _{1}+\cdots +\omega _{k})}\\{} & \displaystyle \hspace{1em}\times \prod \limits_{j=1}^{k}{(\lambda +i\omega _{j})}^{-(\frac{1}{2}-\frac{1-H}{k})}\widehat{B}(d\omega _{1})\dots \widehat{B}(d\omega _{k}),\end{array}\]Proof.
We first observe that
has the Fourier transform
(16)
\[h_{t}(y_{1},\dots ,y_{k})={\int _{0}^{t}}\prod \limits_{j=1}^{k}{(s-y_{j})_{+}^{d-1}}{e}^{-\lambda (s-y_{j})_{+}}\hspace{0.1667em}ds\]
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \widehat{h_{t}}(\omega _{1},\dots ,\omega _{k})\\{} & \displaystyle \hspace{1em}=\frac{1}{{(2\pi )}^{\frac{k}{2}}}\int _{{\mathbb{R}}^{k}}{e}^{i{\textstyle\sum _{j=1}^{k}}\omega _{j}y_{j}}{\int _{0}^{t}}\prod \limits_{j=1}^{k}{(s-y_{j})_{+}^{d-1}}{e}^{-\lambda (s-y_{j})_{+}}\hspace{0.1667em}ds\hspace{0.1667em}dy_{1}\dots dy_{k}\\{} & \displaystyle \hspace{1em}=\frac{1}{{(2\pi )}^{\frac{k}{2}}}\int _{{\mathbb{R}}^{k}}{\int _{0}^{t}}{e}^{i{\textstyle\sum _{j=1}^{k}}\omega _{j}(s-u_{j})}\prod \limits_{j=1}^{k}{(u_{j})_{+}^{d-1}}{e}^{-\lambda (u_{j})_{+}}\hspace{0.1667em}ds\hspace{0.1667em}du_{1}\dots du_{k}\\{} & \displaystyle \hspace{1em}=\frac{1}{{(2\pi )}^{\frac{k}{2}}}{\int _{0}^{t}}\int _{{\mathbb{R}}^{k}}{e}^{is{\textstyle\sum _{j=1}^{k}}\omega _{j}}\prod \limits_{j=1}^{k}{(u_{j})_{+}^{d-1}}{e}^{-(\lambda +i\omega _{j})u_{j}}\hspace{0.1667em}du_{1}\dots du_{k}\hspace{0.1667em}ds\\{} & \displaystyle \hspace{1em}={\bigg[\frac{\varGamma (d)}{\sqrt{2\pi }}\bigg]}^{k}\frac{{e}^{it(\omega _{1}+\cdots +\omega _{k})}-1}{i(\omega _{1}+\cdots +\omega _{k})}\prod \limits_{j=1}^{k}{(\lambda +i\omega _{j})}^{-d}\\{} & \displaystyle \hspace{1em}={\bigg[\frac{\varGamma (\frac{1}{2}-\frac{1-H}{k})}{\sqrt{2\pi }}\bigg]}^{k}\frac{{e}^{it(\omega _{1}+\cdots +\omega _{k})}-1}{i(\omega _{1}+\cdots +\omega _{k})}\prod \limits_{j=1}^{k}{(\lambda +i\omega _{j})}^{-(\frac{1}{2}-\frac{1-H}{k})},\end{array}\]
using the well-known formula for the characteristic function of the gamma density. Then (2), together with Proposition 9.3.1 in [21], implies that
\[\begin{array}{r@{\hskip0pt}l}\displaystyle {Z_{H,\lambda }^{k}}(t)& \displaystyle ={\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}h_{t}(y_{1},\dots ,y_{k})B(dy_{1})\dots B(dy_{k})\\{} & \displaystyle \triangleq {\int _{{\mathbb{R}}^{k}}^{{^{\prime\prime }}}}\widehat{h_{t}}(\omega _{1},\dots ,\omega _{k})\widehat{B}(d\omega _{1})\dots \widehat{B}(d\omega _{k})\\{} & \displaystyle =C_{H,k}{\int _{{\mathbb{R}}^{k}}^{{^{\prime\prime }}}}\frac{{e}^{it(\omega _{1}+\cdots +\omega _{k})}-1}{i(\omega _{1}+\cdots +\omega _{k})}\prod \limits_{j=1}^{k}{(\lambda +i\omega _{j})}^{-(\frac{1}{2}-\frac{1-H}{k})}\widehat{B}(d\omega _{1})\dots \widehat{B}(d\omega _{k}),\end{array}\]
which is equivalent to (15). □3 Limit theorem
In this section, we show that the process ${Z_{H,\lambda }^{k}}(t)$ is the weak convergence limit of a certain discrete chaos process. Our approach follows the seminal work of Bai and Taqqu [3]. When $k=1$ and $\lambda >0$, the discrete process ${Y}^{\lambda ,k}(n)$, (18), is a time series that is useful to model turbulence [20, 24]. When $k=1$ and $\lambda =0$, Davydov [7] (see also Giraitis et al. [12, p. 276] and Whitt [27, Theorem 4.6.1]) established the corresponding invariance principle for ${Y}^{\lambda ,k}(n)$, where the limit involves a fractional Brownian motion. When $k>1$ and $\lambda =0$, Taqqu [26] showed that the weak convergence limit of ${Y}^{\lambda ,k}(n)$ is the Hermite process (1).
The following proposition gives a powerful tool for proving the result of this section.
Proposition 5.
Let
for $N=1,2,\dots $, where $g_{N}\in {L}^{2}({\mathbb{Z}}^{k})$ for $k\ge 1$, and $\{\varepsilon _{n}\}$ is an i.i.d. sequence with mean zero and variance 1. Assume that, for some $f\in {L}^{2}({\mathbb{R}}^{k})$,
(17)
\[Q_{k}(g_{N}):=\sum \limits_{(j_{1},\dots ,j_{k})\in {\mathbb{Z}}^{k}}^{{^{\prime }}}g_{N}(j_{1},\dots ,j_{k})\varepsilon _{j_{1}}\dots \varepsilon _{j_{k}}\]
\[\int _{{\mathbb{R}}^{k}}{\big|\tilde{g}_{N}(u_{1},\dots ,u_{k})-f(u_{1},\dots ,u_{k})\big|}^{2}\hspace{0.1667em}du_{1}\dots du_{k}\to 0,\hspace{1em}\textit{as }N\to \infty ,\]
where
\[\tilde{g}_{N}(u_{1},\dots ,u_{k}):={N}^{\frac{k}{2}}g_{N}\big([u_{1}N]+c_{1},\dots ,[u_{k}N]+c_{k}\big),\hspace{1em}(c_{1},\dots ,c_{k})\in {\mathbb{Z}}^{k}.\]
Then
\[Q_{k}(g_{N})\stackrel{f.d.d.}{\longrightarrow }\int _{{\mathbb{R}}^{k}}f(u_{1},\dots ,u_{k})B(du_{1})\dots B(du_{k})\]
as $N\to \infty $.
Define the discrete chaos process
where the prime ${^{\prime }}$ indicates exclusion of the diagonals $i_{p}=i_{q}$, $p\ne q$, $\{\varepsilon _{n}\}$ is as before, and
for $d\in (\frac{1}{2}-\frac{1}{2k},\infty )$ and $\lambda >0$. Now, consider
(18)
\[{Y}^{\lambda ,k}(n):=\sum \limits_{(i_{1},i_{2},\dots ,i_{k})\in {\mathbb{Z}}^{k}}^{{^{\prime }}}{C}^{\lambda }(i_{1},i_{2},\dots ,i_{k})\varepsilon _{n-i_{1}}\dots \varepsilon _{n-i_{k}},\](19)
\[{C}^{\lambda }(i_{1},i_{2},\dots ,i_{k})=\prod \limits_{j=1}^{k}{(i_{j})_{+}^{d-1}}{e}^{-\lambda (i_{j})_{+}}\]Theorem 1.
Remark 1.
The Lamperti’s theorem [15] states that if
and $d(N)\to \infty $ as $N\to \infty $, where $\{Y_{k}\}$ is stationary, then $\{Z(t)\}_{t\ge 0}$ is self-similar with stationary increments ($\stackrel{f.d.d.}{\longrightarrow }$ means the convergence of finite-dimensional distributions). In our case, since the stationary processes $\{{Y_{k}^{\frac{\lambda }{N}}}\}$ depend on N through the parameter λ, the limit process $\{{Z_{H,\lambda }^{k}}(t)\}_{t\ge 0}$ need not be a self-similar process. Therefore, the result of Theorem 1 does not contradict the Lamperti theorem.
Proof of Theorem 1.
First, we show that
where
as $N\to \infty $, where
(21)
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \frac{1}{{N}^{H}}{S_{N}^{\frac{\lambda }{N}}}(t)& \displaystyle =\frac{1}{{N}^{H}}\sum \limits_{n=1}^{[Nt]}{Y}^{\frac{\lambda }{N},k}(n)\\{} & \displaystyle =\sum \limits_{(i_{1},\dots ,i_{k})\in {\mathbb{Z}}^{k}}\frac{1}{{N}^{H}}\sum \limits_{n=1}^{[Nt]}{C}^{\frac{\lambda }{N}}(n-i_{1},\dots ,n-i_{k})\varepsilon _{i_{1}}\dots \varepsilon _{i_{k}}\\{} & \displaystyle =Q_{k}(h_{t,N})\stackrel{f.d.d.}{\longrightarrow }{Z_{H,\lambda }^{k}}(t)\hspace{1em}\text{as}\hspace{2.5pt}N\to \infty ,\end{array}\]
\[h_{t,N}(i_{1},\dots ,i_{k}):=\frac{1}{{N}^{H}}\sum \limits_{n=1}^{[Nt]}{C}^{\frac{\lambda }{N}}(n-i_{1},\dots ,n-i_{k}),\]
and $Q_{k}(\cdot )$ is defined by (17). In order to show (21), we just need to check that
(22)
\[\big\| \tilde{h}_{t,N}(y_{1},\dots ,y_{k})-h_{t}(y_{1},\dots ,y_{k})\big\| _{{L}^{2}({\mathbb{R}}^{k})}\to 0\]
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \tilde{h}_{t,N}(y_{1},\dots ,y_{k})& \displaystyle :={N}^{\frac{k}{2}}h_{t,N}\big([Ny_{1}]+1,\dots ,[Ny_{k}]+1\big)\\{} & \displaystyle =\frac{{N}^{\frac{k}{2}}}{{N}^{H}}\sum \limits_{n=1}^{[Nt]}{C}^{\frac{\lambda }{N}}\big(n-[Ny_{1}]-1,\dots ,n-[Ny_{k}]-1\big),\end{array}\]
and $h_{t}(y_{1},\dots ,y_{k})$ is given by (3). Write
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \tilde{h}_{t,N}(y_{1},\dots ,y_{k})& \displaystyle =\frac{{N}^{\frac{k}{2}}}{{N}^{H}}\sum \limits_{n=1}^{[Nt]}{C}^{\frac{\lambda }{N}}\big(n-[Ny_{1}]-1,\dots ,n-[Ny_{k}]-1\big)\\{} & \displaystyle =\frac{1}{{N}^{1+kd-k}}\sum \limits_{n=1}^{[Nt]}\prod \limits_{i=1}^{k}{\big(n-[Ny_{i}]-1\big)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}(n-[Ny_{i}]-1)_{+}}\\{} & \displaystyle =\frac{1}{N}\sum \limits_{n=1}^{[Nt]}\prod \limits_{i=1}^{k}{\bigg(\frac{n-[Ny_{i}]-1}{N}\bigg)_{+}^{d-1}}{e}^{-\lambda (\frac{n-[Ny_{i}]-1}{N})_{+}}\\{} & \displaystyle ={\int _{0}^{t}}\prod \limits_{i=1}^{k}{\bigg(\frac{[Ns]-[Ny_{i}]}{N}\bigg)_{+}^{d-1}}{e}^{-\lambda (\frac{[Ns]-[Ny_{i}]}{N})_{+}}\hspace{0.1667em}ds.\end{array}\]
Let $d=1$. In this case,
\[{\bigg(\frac{[Ns]-[Ny]}{N}\bigg)_{+}^{d-1}}{e}^{-\lambda (\frac{[Ns]-[Ny]}{N})_{+}}={e}^{-\lambda (\frac{[Ns]-[Ny]}{N})_{+}}\le {e}^{-\lambda (s-y)_{+}}{e}^{\frac{\lambda }{N}}\]
for all $N\ge 1$, and hence
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \Bigg|\prod \limits_{i=1}^{k}{e}^{-\lambda (\frac{[Ns]-[Ny_{i}]}{N})_{+}}\Bigg|& \displaystyle \le {e}^{\lambda k}\prod \limits_{i=1}^{k}{e}^{-\lambda (s-y_{i})_{+}}\\{} & \displaystyle =:g_{1}(s-y_{1},\dots ,s-y_{k}).\end{array}\]
Next, consider $0<d<1$. Since $[Ns]-[Ny]>Ns-Ny-1$, we get
\[{\bigg(\frac{[Ns]-[Ny]}{N}\bigg)_{+}^{d-1}}<{\bigg(\frac{Ns-Ny-1}{N}\bigg)_{+}^{d-1}}\le {(s-y-1)_{+}^{d-1}}\]
for all $N\ge 1$, and hence
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \Bigg|\prod \limits_{i=1}^{k}{\bigg(\frac{[Ns]-[Ny_{i}]}{N}\bigg)_{+}^{d-1}}{e}^{-\lambda (\frac{[Ns]-[Ny_{i}]}{N})_{+}}\Bigg|& \displaystyle <\prod \limits_{i=1}^{k}{(s-y_{i}-1)_{+}^{d-1}}{e}^{-\lambda (s-y_{i}-1)_{+}}\\{} & \displaystyle =:g_{2}(s-y_{1},\dots ,s-y_{k}).\end{array}\]
Finally, suppose that $d>1$. Since $[Ns]-[Ny]<Ns-Ny+1$, we get
\[{\bigg(\frac{[Ns]-[Ny]}{N}\bigg)_{+}^{d-1}}<{\bigg(\frac{Ns-Ny+1}{N}\bigg)_{+}^{d-1}}\le {(s-y+1)_{+}^{d-1}}\]
for all $N\ge 1$, and hence
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \Bigg|\prod \limits_{i=1}^{k}{\bigg(\frac{[Ns]-[Ny_{i}]}{N}\bigg)_{+}^{d-1}}{e}^{-\lambda (\frac{[Ns]-[Ny_{i}]}{N})_{+}}\Bigg|& \displaystyle <\prod \limits_{i=1}^{k}{(s-y_{i}+1)_{+}^{d-1}}{e}^{-\lambda (s-y_{i}-1)_{+}}\\{} & \displaystyle =:g_{3}(s-y_{1},\dots ,s-y_{k}).\end{array}\]
By the similar argument of Lemma 1, we can verify that
\[{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg({\int _{0}^{t}}g_{i}(s-y_{1},\dots ,s-y_{k})\hspace{0.1667em}ds\Bigg)}^{2}\hspace{0.1667em}dy_{1},\dots ,dy_{k}<\infty \]
for $i=1,2,3$. On the other hand, since ${C}^{\lambda }(i_{1},\dots ,i_{k})$ is continuous a.e., ${C}^{\lambda }(\frac{[Ns]-[Ny_{1}]}{N},\dots ,\frac{[Ns]-[Ny_{k}]}{N})$ converges a.e. to ${C}^{\lambda }(s-y_{1},\dots ,s-y_{k})$ as $N\to \infty $. Now apply the dominated convergence theorem to get the desired result (22).In order to show the tightness, we need to verify that
where $\gamma >0$ and $\alpha >\frac{1}{2}$ (here $\{F_{n}\}_{n\ge 1}$ is a sequence of nondecreasing continuous functions on $[0,1]$ that are uniformly bounded and satisfy
where $\omega _{\delta }(F):=\sup _{|t-s|<\delta }|F(t)-F(s)|$ for $\delta >0$). See Lemma 4.4.1 in [12] for more details. Observe that
(23)
\[\mathbb{E}{\big|{N}^{-H}\big({S_{N}^{\frac{\lambda }{N}}}(t)-{S_{N}^{\frac{\lambda }{N}}}(s)\big)\big|}^{2\gamma }\le C{\big|F_{n}(t)-F_{n}(s)\big|}^{2\alpha },\hspace{1em}0\le s<t\le 1,\]
\[\begin{array}{r@{\hskip0pt}l}\displaystyle {S_{N}^{\frac{\lambda }{N}}}(t)& \displaystyle =\sum \limits_{n=1}^{[Nt]}{Y}^{\frac{\lambda }{N},k}(n)=\sum \limits_{(i_{1},\dots ,i_{k})\in \mathbb{Z}}^{{^{\prime }}}\sum \limits_{n=1}^{[Nt]}{C}^{\frac{\lambda }{N},k}(n-i_{1},\dots ,n-i_{k})\varepsilon _{i_{1}}\dots \varepsilon _{i_{k}}\\{} & \displaystyle =\sum \limits_{(i_{1},\dots ,i_{k})\in \mathbb{Z}}^{{^{\prime }}}\sum \limits_{n=1}^{[Nt]}\prod \limits_{j=1}^{k}{(n-i_{j})_{+}^{d-1}}{e}^{-\frac{\lambda }{N}(n-i_{j})_{+}}\varepsilon _{i_{1}}\dots \varepsilon _{i_{k}}\\{} & \displaystyle =\sum \limits_{(i_{1},\dots ,i_{k})\in \mathbb{Z}}^{{^{\prime }}}{N}^{kd-k+1}\Bigg[\frac{1}{N}\sum \limits_{n=1}^{[Nt]}\prod \limits_{j=1}^{k}{\bigg(\frac{n-i_{j}}{N}\bigg)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}(n-i_{j})_{+}}\Bigg]\varepsilon _{i_{1}}\dots \varepsilon _{i_{k}}\\{} & \displaystyle =\sum \limits_{(i_{1},\dots ,i_{k})\in \mathbb{Z}}^{{^{\prime }}}{N}^{kd-k+1}\Bigg[{\int _{0}^{t}}\prod \limits_{j=1}^{k}{\bigg(\frac{[Ny]+1-i_{j}}{N}\bigg)_{+}^{d-1}}\\{} & \displaystyle \hspace{1em}\times {e}^{-\frac{\lambda }{N}([Ny]+1-i_{j})_{+}}\hspace{0.1667em}dy\Bigg]\varepsilon _{i_{1}}\dots \varepsilon _{i_{k}}\\{} & \displaystyle =N\sum \limits_{(i_{1},\dots ,i_{k})\in \mathbb{Z}}^{{^{\prime }}}\Bigg[{\int _{0}^{t}}\prod \limits_{j=1}^{k}{\big([Ny]+1-i_{j}\big)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}([Ny]+1-i_{j})_{+}}\hspace{0.1667em}dy\Bigg]\varepsilon _{i_{1}}\dots \varepsilon _{i_{k}}.\end{array}\]
Therefore, we get
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \mathbb{E}{\big|{N}^{-H}\big({S_{N}^{\frac{\lambda }{N}}}(t)\hspace{0.1667em}-\hspace{0.1667em}{S_{N}^{\frac{\lambda }{N}}}(s)\big)\big|}^{2}\\{} & \displaystyle \hspace{1em}={N}^{2-2H}\\{} & \displaystyle \hspace{2em}\times \mathbb{E}{\Bigg|\sum \limits_{(i_{1},\dots ,i_{k})\in \mathbb{Z}}^{{^{\prime }}}\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{\big([Ny]+1-i_{j}\big)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}([Ny]+1-i_{j})_{+}}\hspace{0.1667em}dy\Bigg]\varepsilon _{i_{1}}\dots \varepsilon _{i_{k}}\Bigg|}^{2}\\{} & \displaystyle \hspace{1em}\le k!{N}^{2-2H}\sum \limits_{(i_{1},\dots ,i_{k})\in \mathbb{Z}}^{{^{\prime }}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{\big([Ny]+1-i_{j}\big)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}([Ny]+1-i_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\\{} & \displaystyle \hspace{1em}=k!{N}^{2-2H+k}\\{} & \displaystyle \hspace{2em}\times {\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{\big([Ny]+1-[Nx_{j}]\big)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}([Ny]+1-[Nx_{j}])_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dx_{1}\dots dx_{k}.\end{array}\]
Now, we consider two different cases corresponding with the range of d. First, assume that $\frac{1}{2}-\frac{1}{2k}<d\le 1$ (equivalently, $\frac{1}{2}<H\le 1+\frac{k}{2}$): Since $[Ny]-[Nx_{j}]+1>Ny-Nx_{j}$, we can write
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \mathbb{E}{\big|{N}^{-H}\big({S_{N}^{\frac{\lambda }{N}}}(t)-{S_{N}^{\frac{\lambda }{N}}}(s)\big)\big|}^{2}\\{} & \displaystyle \hspace{1em}\le k!{N}^{2-2H+k}\\{} & \displaystyle \hspace{2em}\times {\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{\big([Ny]+1-[Nx_{j}]\big)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}([Ny]+1-[Nx_{j}])_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dx_{1}\dots dx_{k}\\{} & \displaystyle \hspace{1em}\le k!{N}^{2-2H+2kd-k}{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-x_{j})_{+}^{d-1}}{e}^{-\lambda (y-x_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dx_{1}\dots dx_{k}\\{} & \displaystyle \hspace{1em}=k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-x_{j})_{+}^{d-1}}{e}^{-\lambda (y-x_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dx_{1}\dots dx_{k}.\end{array}\]
Now, let $d>1$. Since $Ny-Nx_{j}<[Ny]-[Nx_{j}]+1<Ny-Nx_{j}+N$, we have
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \mathbb{E}{\big|{N}^{-H}\big({S_{N}^{\frac{\lambda }{N}}}(t)-{S_{N}^{\frac{\lambda }{N}}}(s)\big)\big|}^{2}\\{} & \displaystyle \hspace{1em}\le k!{N}^{2-2H+k}\\{} & \displaystyle \hspace{2em}\times {\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{\big([Ny]+1-[Nx_{j}]\big)_{+}^{d-1}}{e}^{-\frac{\lambda }{N}([Ny]+1-[Nx_{j}])_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dx_{1}\dots dx_{k}\\{} & \displaystyle \hspace{1em}\le k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-x_{j}+1)_{+}^{d-1}}{e}^{-\lambda (y-x_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dx_{1}\dots dx_{k}\\{} & \displaystyle \hspace{1em}=k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-x_{j}+1)}^{d-1}{e}^{-\lambda (y-x_{j}+1)}{e}^{\lambda }\mathbf{1}_{\{y>x_{j}\}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dx_{1}\dots dx_{k}\\{} & \displaystyle \hspace{1em}={e}^{2\lambda k}k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-z_{j})}^{d-1}{e}^{-\lambda (y-z_{j})}\mathbf{1}_{\{y>z_{j}+1\}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dz_{1}\dots dz_{k}\\{} & \displaystyle \hspace{1em}\le {e}^{2\lambda k}k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-z_{j})_{+}^{d-1}}{e}^{-\lambda (y-z_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dz_{1}\dots dz_{k}.\end{array}\]
Therefore,
\[\begin{array}{r@{\hskip0pt}l}& \displaystyle \mathbb{E}{\big|{N}^{-H}\big({S_{N}^{\frac{\lambda }{N}}}(t)-{S_{N}^{\frac{\lambda }{N}}}(s)\big)\big|}^{2}\\{} & \displaystyle \hspace{1em}\le {e}^{2\lambda k}k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-z_{j})_{+}^{d-1}}{e}^{-\lambda (y-z_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dz_{1}\dots dz_{k}\end{array}\]
for any $d>\frac{1}{2}-\frac{1}{2k}$ (equivalently, $H>\frac{1}{2}$). According to the proof of Lemma 1,
\[{e}^{2\lambda k}k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-z_{j})_{+}^{d-1}}{e}^{-\lambda (y-z_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dz_{1}\dots dz_{k}\le C|t-s{|}^{2H}\]
for $\frac{1}{2}<H<1$ and
\[{e}^{2\lambda k}k!{\int _{{\mathbb{R}}^{k}}^{{^{\prime }}}}{\Bigg[{\int _{0}^{t-s}}\prod \limits_{j=1}^{k}{(y-z_{j})_{+}^{d-1}}{e}^{-\lambda (y-z_{j})_{+}}\hspace{0.1667em}dy\Bigg]}^{2}\hspace{0.1667em}dz_{1}\dots dz_{k}\le C|t-s{|}^{2}\]
for $H>1$. Now, it remains to apply (23) by selecting $\gamma =1$, $\alpha =\min \{H,1\}$, and $F_{n}(t)=t$ to get the desired result. □