1 Introduction
We consider a family $\{{X_{t}^{x}},t\ge 0,x\in \mathbb{R}\}$ of one-dimensional homogeneous diffusion processes defined on a complete filtered probability space $\{\varOmega ,\mathcal{F},\{\mathcal{F}_{t}\}_{t\ge 0},\mathsf{P}\}$ by a stochastic differential equation
with initial condition ${X_{0}^{x}}=x\in \mathbb{R}$, where $\{W_{t},t\ge 0\}$ is a standard $\mathcal{F}_{t}$-Wiener process. If the initial condition is not important, we will denote the process in question by X. Let the coefficients $a,b$ of equation (1) be continuous and satisfy any conditions of the existence of a nonexplosive weak solution on $\mathbb{R}$. Assume also that $a(x)\ne 0$ for $x\in \mathbb{R}$. We further introduce several objects related to the family $\{{X_{t}^{x}},t\ge 0,x\in \mathbb{R}\}$.
The generator of a diffusion process X is defined for $f\in {C}^{2}(\mathbb{R})$ as
Define the functions
\[ \varphi (x_{0},x)=\exp \Bigg\{-2{\int _{x_{0}}^{x}}\frac{b(u)}{a{(u)}^{2}}du\Bigg\},\hspace{2em}\varPhi (x_{0},x)={\int _{x_{0}}^{x}}\varphi (x_{0},z)dz,\]
$x_{0},x\in \mathbb{R}\cup \{-\infty ,+\infty \}$. It is easy to see that, for a fixed $x_{0}\in \mathbb{R}$, the function $\varPhi (x_{0},\cdot )$ solves a second-order homogeneous differential equation $\mathcal{L}\varPhi (x_{0},\cdot )=0$.For $x,y\in \mathbb{R}$, let ${\tau _{y}^{x}}=\inf \{t\ge 0,{X_{t}^{x}}=y\}$ be the first moment of hitting the point y. For any $(a,b)\subset \mathbb{R}$ and $x\in (a,b)$, let ${\tau _{a,b}^{x}}=\inf \{t\ge 0,{X_{t}^{x}}\notin (a,b)\}={\tau _{a}^{x}}\wedge {\tau _{b}^{x}}$ be the first moment of exiting the interval $(a,b)$. (We use the convention $\inf \varnothing =+\infty $.)
For any $t>0$ and $y\in \mathbb{R}$, define the local time of the process ${X}^{x}$ at the point y on the interval $[0,t]$ by
(The factor ${a}^{2}(y)$ is included to agree with the general Meyer–Tanaka definition of the local time of a semimartingale [7].) The limit in (2) exists almost surely and defines a continuous nondecreasing process $\{{L_{t}^{x}}(y),t\ge 0\}$ for any $x,y\in \mathbb{R}$. The local time on the whole interval $[0,+\infty )$ will be denoted by ${L_{\infty }^{x}}(y)=\lim _{t\to +\infty }{L_{t}^{x}}(y)$.
(2)
\[ {L_{t}^{x}}(y)=a{(y)}^{2}\underset{\varepsilon \downarrow 0}{\lim }\frac{1}{2\varepsilon }{\int _{0}^{t}}\mathbb{I}\big\{\big|{X_{s}^{x}}-y\big|\le \varepsilon \big\}ds.\]In this article, we focus on the transient diffusion processes, that is, those converging to $+\infty $ or $-\infty $ as $t\to \infty $. We use the explicit distribution of ${L_{\infty }^{x}}$ to study integral functionals of the form $J_{\infty }(f)={\int _{0}^{\infty }}f({X_{s}^{x}})ds$, which can be interpreted as continuous perpetuities in the framework of financial mathematics. We follow the approach of Salminen and Yor [8] to study integral functionals of a Wiener process with positive drift and generalize their results to homogeneous transient diffusion processes. Applying the results of [6], we establish criteria of convergence of almost sure finiteness of the functionals $J_{\infty }(f)$, calculate their moments and potentials, and bound their exponential moments.
2 The distribution of the local time of a transient diffusion process
In this section, we concentrate on the explicit distribution of ${L_{\infty }^{x}}(y)$. According to the classical results (see, e.g., [4]), in the case where $\varPhi (x,+\infty )=-\varPhi (x,-\infty )=+\infty $ for some (equivalently, for all) $x\in \mathbb{R}$, the diffusion process X is recurrent, that is,
\[ \mathsf{P}\Big(\underset{t\to +\infty }{\limsup }{X_{t}^{x}}=+\infty ,\underset{t\to +\infty }{\liminf }{X_{t}^{x}}=-\infty \Big)=1.\]
Therefore, ${L_{\infty }^{x}}(y)=+\infty $ for all $x,y\in \mathbb{R}$ a.s.In what follows, we will consider only the case of a transient process X, where at least one of the integrals $\varPhi (x,+\infty )$ and $\varPhi (x,-\infty )$ is finite. We formulate the following statement concerning the distribution of ${L_{\infty }^{x}}(y)$ that can be easily deduced from the results of [3, 1].
Proposition 1.
1. In each of the cases $x=y$, $x<y$ and $-\varPhi (0,-\infty )=+\infty $, and $x>y$ and $\varPhi (0,+\infty )=+\infty $, the local time ${L_{\infty }^{x}}(y)$ is exponentially distributed with parameter $\psi _{y}(0)$ given by (5).
2. If $x<y$ and $-\varPhi (0,-\infty )<+\infty $, then the local time ${L_{\infty }^{x}}(y)$ is distributed as $\kappa \xi $, where ξ is exponentially distributed with parameter $\psi _{y}(0)$, and κ is an independent of ξ Bernoulli random variable with
3. If $x>y$ and $\varPhi (0,+\infty )<+\infty $, then the local time ${L_{\infty }^{x}}(y)$ is distributed as $\kappa \xi $, where ξ is exponentially distributed with parameter $\psi _{y}(0)$, and κ is an independent of ξ Bernoulli random variable with
Proof.
By the strong Markov property of the process X, for any $l\ge 0$ and $x,y\in \mathbb{R}$,
For $x<y$,
Thus, it is sufficient to determine the distribution of variables ${L_{\infty }^{x}}(x)$. But it was proved in [1, II.13, II.27] that $\mathsf{P}({L_{\infty }^{x}}(x)>l)=\exp (-l\psi _{x}(0))$, where
with $\frac{1}{\infty }:=0$. Hence, the proof follows. □
\[ \mathsf{P}\big({L_{\infty }^{x}}(y)>l\big)=\mathsf{P}\big({L_{\infty }^{y}}(y)>l\big)\mathsf{P}\big({\tau _{y}^{x}}<+\infty \big).\]
The probability $\mathsf{P}({\tau _{y}^{x}}<+\infty )=1-\mathsf{P}({\tau _{y}^{x}}=+\infty )$ can be found with the help of the well-known formula (see, e.g., [3, Section VIII.6, (18)]): for $x\in (a,b)$,
Then the value of probability in question depends on x, y and on the integrals $\varPhi (x,+\infty )$, $\varPhi (x,-\infty )$. Specifically, if $x>y$, then
\[ \mathsf{P}\big({\tau _{y}^{x}}=\infty \big)=\underset{a\to +\infty }{\lim }\mathsf{P}\big({X_{\tau _{y,a}}^{x}}=a\big)=\underset{a\to +\infty }{\lim }\frac{\varPhi (y,x)}{\varPhi (y,a)},\]
whence
(3)
\[ \mathsf{P}\big({\tau _{y}^{x}}=+\infty \big)=\left\{\begin{array}{l@{\hskip10.0pt}l}\frac{\varPhi (y,x)}{\varPhi (y,+\infty )},\hspace{1em}& \varPhi (x,+\infty )<+\infty ,\\{} 0,\hspace{1em}& \varPhi (x,+\infty )=+\infty .\end{array}\right.\]
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \mathsf{P}\big({\tau _{y}^{x}}=\infty \big)& \displaystyle =\underset{a\to -\infty }{\lim }\big(1-\mathsf{P}\big({X_{\tau _{a,y}}^{x}}=y\big)\big)=\underset{a\to -\infty }{\lim }\frac{\varPhi (a,y)-\varPhi (a,x)}{\varPhi (a,y)}\\{} & \displaystyle =\underset{a\to -\infty }{\lim }\frac{\varphi (a,x)\varPhi (x,y)}{-\varphi (a,y)\varPhi (y,a)}=\underset{a\to -\infty }{\lim }\frac{-\varphi (a,x)\varphi (x,y)\varPhi (y,x)}{-\varphi (a,y)\varPhi (y,a)}\\{} & \displaystyle =\underset{a\to -\infty }{\lim }\frac{\varPhi (y,x)}{\varPhi (y,a)};\end{array}\]
therefore,
(4)
\[ \mathsf{P}\big({\tau _{y}^{x}}=+\infty \big)=\left\{\begin{array}{l@{\hskip10.0pt}l}\frac{\varPhi (y,x)}{\varPhi (y,-\infty )},\hspace{1em}& -\varPhi (x,-\infty )<+\infty ,\\{} 0,\hspace{1em}& -\varPhi (x,-\infty )=+\infty .\end{array}\right.\](5)
\[ \psi _{x}(0)=\frac{1}{2}\bigg(\frac{1}{\varPhi (x,+\infty )}-\frac{1}{\varPhi (x,-\infty )}\bigg)\]Consider two examples where the parameters of the distribution of local time can be calculated explicitly.
Example 1.
Let $a(x)\hspace{0.1667em}\equiv \hspace{0.1667em}a\hspace{0.1667em}\ne \hspace{0.1667em}0$ and $b(x)\hspace{0.1667em}\equiv \hspace{0.1667em}b$ be constant. Then
and
\[ \varPhi (x,y)=\frac{{a}^{2}}{2b}\big(1-{e}^{-2b(y-x)/{a}^{2}}\big)\hspace{1em}\text{for}\hspace{2.5pt}b\ne 0;\hspace{2em}\varPhi (x,y)=y-x\hspace{1em}\text{for}\hspace{2.5pt}b=0.\]
In this case, the diffusion process X is transient if and only if $b\ne 0$; moreover, $-\varPhi (0,-\infty )=+\infty $ and $\varPhi (0,+\infty )<+\infty $ for $b>0$, and $-\varPhi (0,-\infty )<+\infty $ and $\varPhi (0,+\infty )=+\infty $ for $b<0$. The cases are symmetric; therefore, we consider only the case $b>0$.Now $\psi _{x}(0)=\frac{b}{{a}^{2}}$. Thus, for $x\le y$, the local time ${L_{\infty }^{x}}(y)$ is exponentially distributed with parameter $\frac{b}{{a}^{2}}$. For $x>y$, the local time is distributed as $\kappa \xi $, where ξ has an exponential distribution with parameter $\frac{b}{{a}^{2}}$, and κ is a Bernoulli random variable independent of ξ and distributed as $\mathsf{P}(\kappa =1)=1-\mathsf{P}(\kappa =0)={e}^{-2b(x-y)/{a}^{2}}$. Using the properties of exponential distribution, we see that these cases can be combined: ${L_{\infty }^{x}}(y)\stackrel{d}{=}(\xi -2(x-y)_{+})_{+}$, where $a_{+}=a\vee 0$.
Example 2.
Let $a(x)=\sqrt{{x}^{2}+1}$ and $b(x)=x$. Then $\varphi (x,y)=\frac{{x}^{2}+1}{{y}^{2}+1}$ and $\varPhi (x,y)\hspace{0.1667em}=(1+{x}^{2})(\arctan y-\arctan x)$. We see that the process is transient and $-\varPhi (0,-\infty )=\varPhi (0,\infty )=\frac{\pi }{2}<\infty $.
Due to Corollary 1, the local time ${L_{\infty }^{x}}(y)$ is distributed as $\kappa \xi $, where ξ has an exponential distribution with parameter
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \psi _{x}(0)& \displaystyle =\frac{1}{2\varPhi (x,+\infty )}-\frac{1}{2\varPhi (x,+\infty )}\\{} & \displaystyle =\frac{1}{(1+{x}^{2})(\pi -2\arctan x)}-\frac{1}{(1+{x}^{2})(\pi +2\arctan x)}\\{} & \displaystyle =\frac{4\arctan x}{(1+{x}^{2})({\pi }^{2}-4{\arctan }^{2}x)},\end{array}\]
and κ is a Bernoulli random variable, which is independent of ξ and distributed as
3 Integral functionals of a transient diffusion processes
For a measurable function $f:\mathbb{R}\to \mathbb{R}$ such that f and $f/{a}^{2}$ are locally integrable, define the integral functional
We will study the questions of finiteness and existence of moments of ${J_{\infty }^{x}}(f)$. We start with the well-known occupation density formula (see, e.g., [4, 7])
If the process ${X}^{x}$ is recurrent, then ${L_{\infty }^{x}}(y)=\infty $ a.s. for all $y\in \mathbb{R}$, so ${J_{\infty }^{x}}(f)$ is undefined unless f is identically zero. Therefore, we will require that the process X is transient. We recall that this holds iff $\varPhi (0,+\infty )$ of $\varPhi (0,-\infty )$ is finite. Moreover, if $\varPhi (0,+\infty )<+\infty $ and $\varPhi (0,-\infty )=-\infty $, then ${X_{s}^{x}}\to +\infty $ a.s.; if $\varPhi (0,+\infty )=+\infty $ and $\varPhi (0,-\infty )>-\infty $, then ${X_{s}^{x}}\to -\infty $ a.s.; if $\varPhi (0,+\infty )<+\infty $ and $\varPhi (0,-\infty )>-\infty $, then ${X_{s}^{x}}\to +\infty $ on a set $A_{+}$ of positive probability, and ${X_{s}^{x}}\to -\infty $ on a set $A_{-}=\varOmega \setminus A_{+}$ of positive probability.
We start with a criterion of almost sure finiteness of ${J_{\infty }^{x}}(f)$. It was obtained in [5] in the case where only one of the integrals $\varPhi (0,+\infty )$ and $\varPhi (0,-\infty )$ is finite; a complete analysis was made in [6]. Define
Theorem 1 ([6]).
For arbitrary $x\in \mathbb{R}$, the following statements hold.
-
• Let $\varPhi (0,+\infty )<+\infty $ and $\varPhi (0,-\infty )=-\infty $.
-
• Let $\varPhi (0,+\infty )=+\infty $ and $\varPhi (0,-\infty )>-\infty $.
-
• Let $\varPhi (0,+\infty )<+\infty $ and $\varPhi (0,-\infty )>-\infty $.
-
If $I_{1}(f)<+\infty $, then ${J_{\infty }^{x}}(f)\in \mathbb{R}$ a.s. on $A_{+}$.
-
If $I_{1}(f)=\infty $, then ${J_{\infty }^{x}}(f)=\infty $ a.s. on $A_{+}$.
-
If $I_{2}(f)<+\infty $, then ${J_{\infty }^{x}}(f)\in \mathbb{R}$ a.s. on $A_{-}$.
-
If $I_{2}(f)=+\infty $, then ${J_{\infty }^{x}}(f)=\infty $ a.s. on $A_{-}$.
-
In what follows, we consider the case where $\varPhi (0,+\infty )<+\infty $ and $\varPhi (0,-\infty )=-\infty $, the other cases being similar. The next result is a direct consequence of Proposition 1.
Lemma 1.
Let $\varPhi (0,+\infty )<+\infty $ and $\varPhi (0,-\infty )=-\infty $. Then, for any $k\ge 1$,
\[ \mathsf{E}\big[{L_{\infty }^{x}}{(y)}^{k}\big]=k!{\big(2\varPhi (y,+\infty )\big)}^{k}\hspace{1em}\textit{for}\hspace{2.5pt}x\le y\]
and
\[\begin{array}{r@{\hskip0pt}l}\displaystyle \mathsf{E}\big[{L_{\infty }^{x}}{(y)}^{k}\big]& \displaystyle ={2}^{k}k!\varPhi {(y,+\infty )}^{k-1}\big(\varPhi (y,+\infty )-\varPhi (y,x)\big)\\{} & \displaystyle ={2}^{k}k!\varPhi {(y,+\infty )}^{k-1}\varphi (y,x)\varPhi (x,+\infty )\end{array}\]
for $x>y$.
Example 3.
Let $a=1$ and $b=\mu >0$ with some constant μ, so that X is a Brownian motion with constant positive drift. Furthermore, in this case, $\varPhi (y,+\infty )=1/2\mu $, $\varphi (y,x)=\exp \{-2\mu (x-y)\}$, and $\varPhi (0,-\infty )=-\infty $. Therefore, the criterion for $J_{\infty }(f)$ to be finite is ${\int _{0}^{\infty }}|f(x)|dx<\infty $, which coincides with that of [8]. For what concerns the moments of local times, in this case, $\mathsf{E}[{L_{\infty }^{x}}(y)]=1/\mu $ for $x\le y$ and $\mathsf{E}[{L_{\infty }^{x}}(y)]=\frac{1}{\mu }\exp \{-2\mu (x-y)\}$ for $x>y$.
We further derive conditions for $\mathsf{E}[{J_{\infty }^{x}}(f)]$ to be finite.
Theorem 2.
Let $\varPhi (0,+\infty )<+\infty ,\varPhi (0,-\infty )=-\infty $, and $I_{1}(f)<+\infty $. Assume additionally that
Then
Remark 1.
For a Brownian motion with positive drift μ, a sufficient condition for $\mathsf{E}[{J_{\infty }^{x}}(f)]$ to be finite is
\[ {\int _{x}^{+\infty }}\big|f(u)\big|du+{e}^{-2\mu x}{\int _{-\infty }^{x}}\big|f(u)\big|{e}^{2\mu u}du<\infty ;\]
and in that case, we have the equality
\[ \mathsf{E}\big[{J_{\infty }^{x}}(f)\big]=\frac{1}{\mu }{\int _{x}^{+\infty }}f(u)du+\frac{1}{\mu }{e}^{-2\mu x}{\int _{-\infty }^{x}}f(u){e}^{2\mu u}du.\]
Obviously, the requirement $\int _{\mathbb{R}}|f(u)|du<\infty $ is also sufficient, as stated in [8].Now we continue with the moments of $J_{\infty }(f)$ of higher order.
Theorem 3.
Let $\varPhi (0,+\infty )<+\infty $ and $\varPhi (0,-\infty )=-\infty $. The moments of higher order admit the following bound: for any $k>1$,
(7)
\[\begin{array}{r@{\hskip0pt}l}\displaystyle {\big(\mathsf{E}\big[{\big|{J_{\infty }^{x}}(f)\big|}^{k}\big]\big)}^{1/k}& \displaystyle \le 2{(k!)}^{1/k}\Bigg(\hspace{-0.1667em}{\int _{x}^{+\infty }}\hspace{-0.1667em}\frac{|f(u)|}{a{(u)}^{2}}\varPhi (u,+\infty )du\\{} & \displaystyle \hspace{1em}+\varPhi {(x,+\infty )}^{1/k}\hspace{-0.1667em}{\int _{-\infty }^{x}}\hspace{-0.1667em}\frac{|f(u)|}{a{(u)}^{2}}\varPhi {(u,+\infty )}^{1-1/k}\varphi {(u,x)}^{1/k}du\Bigg).\end{array}\]Proof.
We use representation (6) and the generalized Minkowski inequality to get the following equalities and bounds:
Now (7) follows immediately from (8) and Lemma 1. □
(8)
\[ {\big(\mathsf{E}\big[{\big|{J_{\infty }^{x}}(f)\big|}^{k}\big]\big)}^{1/k}\hspace{-0.1667em}={\bigg(\hspace{-0.1667em}\mathsf{E}\bigg[{\bigg|\int _{\mathbb{R}}\hspace{-0.1667em}\frac{f(y)}{a{(y)}^{2}}{L_{\infty }^{x}}(y)dy\bigg|}^{k}\bigg]\bigg)}^{1/k}\hspace{-0.1667em}\hspace{-0.1667em}\le \hspace{-0.1667em}\int _{\mathbb{R}}\hspace{-0.1667em}\frac{|f(y)|}{a{(y)}^{2}}{\big(\mathsf{E}\big[{L_{\infty }^{x}}{(y)}^{k}\big]\big)}^{1/k}dy.\]We conclude with the existence of potential and exponential moments. Some related results were obtained in [5].
The following result is an immediate corollary of Theorem 2.
Theorem 4.
Let $\varPhi (0,+\infty )<+\infty $, $\varPhi (0,-\infty )=-\infty $, and
\[ P_{0}=2\underset{x\in \mathbb{R}}{\sup }\Bigg({\int _{x}^{+\infty }}\frac{|f(u)|}{a{(u)}^{2}}\varPhi (u,+\infty )du+\varPhi (x,+\infty ){\int _{-\infty }^{x}}\frac{|f(u)|}{a{(u)}^{2}}\varphi (u,x)du\Bigg)<\infty .\]
Then the integral functional $J_{\infty }(f)$ has a bounded potential $P\le P_{0}$.
Proof.
We apply the following result of Dellacherie and Meyer [2], see also [8, Lemma 5.2]. Let A be a continuous adapted nondecreasing process starting at zero such that there exists a constant $C>0$ satisfying $\mathsf{E}[A_{\infty }-A_{t}\mid \mathcal{F}_{t}]\le C$ for any $t\ge 0$. Then
for $\lambda <{C}^{-1}$.