Skip to main content
Log in

Ornstein-Uhlenbeck Processes of Bounded Variation

  • Published:
Methodology and Computing in Applied Probability Aims and scope Submit manuscript

Abstract

Ornstein-Uhlenbeck process of bounded variation is introduced as a solution of an analogue of the Langevin equation with an integrated telegraph process replacing a Brownian motion. There is an interval I such that the process starting from the internal point of I always remains within I. Starting outside, this process a. s. reaches this interval in a finite time. The distribution of the time for which the process falls into this interval is obtained explicitly. The certain formulae for the mean and the variance of this process are obtained on the basis of the joint distribution of the telegraph process and its integrated copy. Under Kac’s rescaling, the limit process is identified as the classical Ornstein-Uhlenbeck process.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  • Andrews GE, Askey R, Roy R (1999) Special functions. Encyclopedia of mathematics and its applications, vol 71. Cambridge University Press

  • Beghin L, Nieddu L, Orsingher E (2001) Probabilistic analysis of the telegrapher’s process with drift by mean of relativistic transformations. J Appl Math Stoch Anal 14:11–25

    Article  MathSciNet  Google Scholar 

  • Buonocore A, Caputo L, D’Onofrio G, Pirozzi E (2015) Closed-form solutions for the first-passage-time problem and neuronal modeling. Ricerche Di Matematica 64 (2):421–439

    Article  MathSciNet  Google Scholar 

  • Cattaneo CR (1958) Sur une forme de l’équation de la chaleur é,liminant le paradoxe d’une propagation instantanée. Comptes Rendus 247(4):431–433

    MathSciNet  MATH  Google Scholar 

  • Coffey WT, Kalmykov YuP, Waldron JT (2004) The Langevin equation: with applications to stochastic problems in physics, chemistry and electrical engineering, 2nd edn. World Scientific Publ

  • Di Crescenzo A, Martinucci B, Zacks S (2014) On the geometric Brownian motion with alternating trend. In: Perna C., Sibillo M. (eds) Mathematical and statistical methods for actuarial sciences and finance. Springer, pp 81–85

  • Di Crescenzo A, Martinucci B, Zacks S (2018) Telegraph process with elastic boundary at the origin. Methodol Comput Appl Probab 20:333–352

    Article  MathSciNet  Google Scholar 

  • Genadot A (2019) Averaging for some simple constrained Markov processes. Probab Math Stat 39(1):139–158

    Article  MathSciNet  Google Scholar 

  • Gradshteyn IS, Ryzhik IM (1994) Table of integrals. Series and Products Academic Press, Boston

    MATH  Google Scholar 

  • Kac M (1974) A stochastic model related to the telegrapher’s equation. Rocky Mountain J Math 4:497–509. Reprinted from: M. Kac, Some stochastic problems in physics and mathematics. Colloquium lectures in the pure and applied sciences, No. 2, hectographed, Field Research Laboratory, Socony Mobil Oil Company, Dallas, TX, 1956, pp. 102–122

    MathSciNet  MATH  Google Scholar 

  • Kolesnik AD (2012) Moment analysis of the telegraph random process. Bul Acad Ştiinţe Repub Mold Mat 68(1):90–107

    MathSciNet  MATH  Google Scholar 

  • Kolesnik AD, Ratanov N (2013) Telegraph processes and option pricing. Springer, Heidelberg

    Book  Google Scholar 

  • López O, Ratanov N (2012) Kac’s rescaling for jump-telegraph processes. Stat Probab Lett 82:1768–1776

    Article  MathSciNet  Google Scholar 

  • López O, Ratanov N (2014) On the asymmetric telegraph processes. J Appl Prob 51:569–589

    Article  MathSciNet  Google Scholar 

  • Maller RA, Müller G, Szimayer A (2009) Ornstein-Uhlenbeck processes and extensions. In: Mikosch T, Kreiß JP, Davis R, Andersen T (eds) Handbook of financial time series. Springer, Berlin, pp 421–437

  • Orsingher E (1995) Motions with reflecting and absorbing barriers driven by the telegraph equation. Random Oper Stoch Equ 3(1):9–21

    Article  MathSciNet  Google Scholar 

  • Ratanov N (2007) A jump telegraph model for option pricing. Quant Finance 7:575–583

    Article  MathSciNet  Google Scholar 

  • Ratanov N (2019) A two-state neuronal model with alternating exponential excitation. Math Biosci Eng 16(5):3411–3434

    Article  MathSciNet  Google Scholar 

  • Ratanov N (2020) First crossing times of telegraph processes with jumps. Methodol Comput Appl Probab 22:349–370

    Article  MathSciNet  Google Scholar 

  • Ratanov N, Di Crescenzo A, Martinucci B (2019) Piecewise deterministic processes following two alternating patterns. J Appl Prob 56:1006–1019

    Article  MathSciNet  Google Scholar 

  • Vasicek O (1977) An equilibrium characterization of the term structure. J Financ Econ 5(2):177–188

    Article  Google Scholar 

  • Zacks S (2004) Generalized integrated telegraph processes and the distribution of related stopping times. J Appl Prob 41:497–507

    Article  MathSciNet  Google Scholar 

  • Zacks S (2017) Sample path analysis and distributions of boundary crossing times. Lecture notes in mathematics. Springer

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nikita Ratanov.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: The Telegraph Process

Appendix: The Telegraph Process

Let \(({\varOmega }, \mathcal F, \mathcal F_{t}, \mathbb {P})\) be the complete filtered probability space. Consider the adapted telegraph process \(\mathbb {T}(t), t\geq 0,\) with two alternating symmetric velocities a and − a,a > 0, switching with positive intensities λ0 and λ1.

The joint distribution of \(\mathbb {T}(t)\) and ε(t) can be expressed by means of the (generalised) density functions

$$ {p_{i}^{j}}(x, t):=\mathbb{P}\{\mathbb{T}(t)\in\mathrm{d} x, {\varepsilon}(t)=j~|~{\varepsilon}(0)=i\}/\mathrm{d} x,\qquad i, j\in\{0, 1\},\quad t>0. $$

The following formulae seem to be generally known, but for the best of my belief, they have never been published.

Theorem A.1

The density functions \({p_{i}^{j}}(x, t), i, j\in \{0, 1\},\) are given by

$$ \begin{aligned} {p_{0}^{0}}(x, t)&=\mathrm{e}^{-{\lambda}_{0}t}\delta(x-a_{0}t) +\frac{\sqrt{{\lambda}_{0}{\lambda}_{1}}}{a_{0}-a_{1}}\sqrt{\frac{\xi}{t-\xi}} \mathrm{e}^{-{\lambda}_{0}\xi-{\lambda}_{1}(t-\xi)}I_{1}(2\sqrt{{\lambda}_{0}{\lambda}_{1}\xi(t-\xi)})\mathbbm{1}_{\{0<\xi<t\}}, \\ {p_{1}^{1}}(x, t)&= \mathrm{e}^{-{\lambda}_{1}t}\delta(x-a_{1}t) +\frac{\sqrt{{\lambda}_{0}{\lambda}_{1}}}{a_{0}-a_{1}}\sqrt{\frac{t-\xi}{\xi}}\mathrm{e}^{-{\lambda}_{0}\xi-{\lambda}_{1}(t-\xi)}I_{1}(2\sqrt{{\lambda}_{0}{\lambda}_{1}\xi(t-\xi)})\mathbbm{1}_{\{0<\xi<t\}},\\ {p_{0}^{1}}(x, t)&=\frac{\lambda_{0}}{a_{0}-a_{1}}\mathrm{e}^{-\lambda_{0}\xi-\lambda_{1}(t-\xi)}I_{0}(2\sqrt{\lambda_{0}\lambda_{1}\xi(t-\xi)})\mathbbm{1}_{\{0<\xi<t\}},\\ {p_{1}^{0}}(x, t)&=\frac{{\lambda}_{1}}{a_{0}-a_{1}}\mathrm{e}^{-{\lambda}_{0}\xi-{\lambda}_{1}(t-\xi)}I_{0}(2\sqrt{{\lambda}_{0}{\lambda}_{1}\xi(t-\xi)})\mathbbm{1}_{\{0<\xi<t\}}, \end{aligned} $$
(A.1)

where ξ = (xa1t)/(a0a1),tξ = (a0tx)/(a0a1),a1t < x < a0t.

Here I0 and I1 are the modified Bessel functions,

$$ I_{0}(z)=1+\sum\limits_{n=1}^{\infty}\frac{(z/2)^{2n}}{(n!)^{2}},\qquad I_{1}(z)=I_{0}^{\prime}(z)=\sum\limits_{n=1}^{\infty}\frac{(z/2)^{2n-1}}{(n-1)!n!}. $$

Proof

Let N(t) be the number of velocity switchings in the time interval [0,t).

By virtue of Kolesnik and Ratanov (2013, (4.1.10)-(4.1.11)), \({p_{0}^{0}},\) can be represented as

$$ \begin{aligned} {p_{0}^{0}}(x, t) =&\sum\limits_{n=0}^{\infty}\mathbb{P}\{\mathbb{T}(t)\in\mathrm{d} x, N(t)=2n~|~{\varepsilon}(0)=0\}/\mathrm{d} x\\ =\mathrm{e}^{-{\lambda}_{0}t}\delta(x-a_{0}t)& +\frac{\exp(-{\lambda}_{0}\xi-{\lambda}_{1}(t-\xi))}{a_{0}-a_{1}} \sum\limits_{n=1}^{\infty}\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(n-1)!n!}\xi^{n}(t-\xi)^{n-1}\mathbbm{1}_{\{0<\xi<t\}}\\ =&\mathrm{e}^{-{\lambda}_{0}t}\delta(x-a_{0}t) +\frac{\sqrt{{\lambda}_{0}{\lambda}_{1}}}{a_{0}-a_{1}}\sqrt{\frac{\xi}{t-\xi}}I_{1}(2\sqrt{{\lambda}_{0}{\lambda}_{1}\xi(t-\xi)})\mathbbm{1}_{\{0<\xi<t\}}. \end{aligned}$$

see Gradshteyn and Ryzhik (1994, formula 8.445). The remaining equalities of Eq. A.1 are obtained in the same manner. □

The well-known formulae for the (conditional) distribution of \(\mathbb {T}(t)\) follow from Eq. A.1:

$$ \begin{aligned} p_{0}(x, t)= \mathbb{P}\{\mathbb{T}(t)\in\mathrm{d} x~|~{\varepsilon}(0)=0\}/\mathrm{d} x &= {p_{0}^{0}}(x, t)+{p_{0}^{1}}(x, t), \\ p_{1}(x, t)= \mathbb{P}\{\mathbb{T}(t)\in\mathrm{d} x~|~{\varepsilon}(0)=1\}/\mathrm{d} x & = {p_{1}^{0}}(x, t)+{p_{1}^{1}}(x, t), \end{aligned} $$

cf Beghin et al. (2001) and López and Ratanov (2014) or see in the book by Kolesnik and Ratanov, (2013, (4.1.15)).

Similarly, \(f(x, t)={p_{0}^{0}}(x, t)+{p_{1}^{0}}(x, t)\) (and \(b(x, t)={p_{0}^{1}}(x, t)+{p_{1}^{1}}(x, t)\)) are the distribution density functions of the moving forward (and backward) particles, cf Orsingher (1995), where these formulae were presented in the symmetric case, λ0 = λ1.

The rest of this section is devoted to a description of the first and the second moments of \(\mathbb {T}(t)\) and \(\mathbb {T}(t)\mathbbm {1}_{\{{\varepsilon }(t)=j\}}, j\in \{0, 1\}.\)

We will use the following notations

$$ \mathbb{E}_{i}[g(\mathbb{T}(t))]=\mathbb{E}[g(\mathbb{T}(t))~|~{\varepsilon}(0)=i]={\int}_{-\infty}^{\infty} g(x)p_{i}(x, t)\mathrm{d} x $$

and

$$ {\mathbb{E}_{i}^{j}}[g(\mathbb{T}(t))]=\mathbb{E}[g(\mathbb{T}(t))\cdot\mathbbm{1}_{\{{\varepsilon}(t)=j\}}~|~{\varepsilon}(0)=i]={\int}_{-\infty}^{\infty} g(x){p_{i}^{j}}(x, t)\mathrm{d} x, \qquad i, j\in\{0, 1\}. $$

Theorem A.2

Let a0 = −a1 = a > 0.

For t ≥ 0

$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{0}^{0}}\mathbb{T}(t)&=&a\mathrm{e}^{-{\lambda}_{0}t}\sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!}t^{2n+1}G_{n}^{(1)}(t), \end{array} $$
(A.2)
$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{0}^{1}}\mathbb{T}(t) & =&a\mathrm{e}^{-{\lambda}_{0}t}\sum\limits_{n=0}^{\infty}\frac{{\lambda}_{0}^{n+1}{{\lambda}_{1}^{n}}}{(2n+1)!}t^{2n+2}H_{n}^{(1)}(t), \end{array} $$
(A.3)
$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{1}^{0}}\mathbb{T}(t)& =&-a\mathrm{e}^{-{\lambda}_{1}t}\sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{\lambda}_{1}^{n+1}}{(2n+1)!}t^{2n+2}H_{n}^{(1)}(-t), \end{array} $$
(A.4)
$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{1}^{1}}\mathbb{T}(t)&=&-a\mathrm{e}^{-{\lambda}_{1}t}\sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!}t^{2n+1}G_{n}^{(1)}(-t), \end{array} $$
(A.5)

and

$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{0}^{0}}\mathbb{T}(t)^{2}&=&a^{2}\exp(-{\lambda}_{0}t) \sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!}t^{2n+2}G_{n}^{(2)}(t), \end{array} $$
(A.6)
$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{0}^{1}} \mathbb{T}(t)^{2}&=&a^{2}\exp(-{\lambda}_{0}t) \sum\limits_{n=0}^{\infty}\frac{{\lambda}_{0}^{n+1}{{\lambda}_{1}^{n}}}{(2n+1)!}t^{2n+3}H_{n}^{(2)}(t), \end{array} $$
(A.7)
$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{1}^{0}}\mathbb{T}(t)^{2}&=&a^{2}\exp(-{\lambda}_{1}t) \sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{\lambda}_{1}^{n+1}}{(2n+1)!}t^{2n+3}H_{n}^{(2)}(-t), \end{array} $$
(A.8)
$$ \begin{array}{@{}rcl@{}} {\mathbb{E}_{1}^{1}} \mathbb{T}(t)^{2}&=&a^{2}\exp(-{\lambda}_{1}t) \sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!}t^{2n+2}G_{n}^{(2)}(-t), \end{array} $$
(A.9)

where

$$ G_{n}^{(1)}(t)=-\frac{2n}{2n+1}{\varPhi}(n+1, 2n+2; 2\upbeta t)+{\varPhi}(n, 2n+1; 2\upbeta t), $$
(A.10)
$$ H_{n}^{(1)}(t)=-{\varPhi}(n+2, 2n+3; 2\upbeta t)+{\varPhi}(n+1, 2n+2; 2\upbeta t) $$
(A.11)

and

$$ G_{n}^{(2)}(t)=\frac{2n}{2n+1}{\varPhi}(n+2, 2n+3; 2\upbeta t) -\frac{4n}{2n+1}{\varPhi}(n+1, 2n+2; 2\upbeta t)+{\varPhi}(n, 2n+1; 2\upbeta t), $$
(A.12)
$$ H_{n}^{(2)}(t)=\frac{2n+4}{2n+3}{\varPhi}(n+3, 2n+4; 2\upbeta t) -2{\varPhi}(n+2, 2n+3; 2\upbeta t)+{\varPhi}(n+1, 2n+2; 2\upbeta t), $$
(A.13)

2β = λ0λ1.

Here Φ(a,b; z) denotes the confluent hypergeometric function,

$$ {\varPhi}(\alpha, \upbeta; z):={\sum}_{n=0}^{\infty}\frac{(\alpha)_{n}}{(\upbeta)_{n}}\frac{z^{n}}{n!}, $$

(⋅)n is the Pochhammer symbol; (γ)n = γ(γ + 1)…(γ + n − 1),n ≥ 1, (γ)0 = 1.

Proof

Consider

$$ \begin{aligned} \psi_{i}(z, t)=&\mathbb{E}_{i}\exp(z\mathbb{T}(t))=\mathbb{E}\left[\exp(z\mathbb{T}(t))~|~{\varepsilon}(0)=i\right],\\ \psi_{i}(z, t; n)=&\mathbb{E}_{i}\left[\exp(z\mathbb{T}(t))\mathbbm{1}_{\{N(t)=n\}}\right],\qquad n\geq0, \end{aligned} \qquad i\in\{0, 1\},$$

and

$$ {\psi_{i}^{j}}(z, t)={\mathbb{E}_{i}^{j}}\left[\exp(z\mathbb{T}(t))\right]=\mathbb{E}_{i}\left[\exp(z\mathbb{T}(t))\mathbbm{1}_{\{{\varepsilon}(t)=j\}}\right],\qquad i, j\in\{0, 1\}, $$

corresponding to the moment generating function of \(\mathbb {T}(t).\) Notice that \(\psi _{i}(z, t)=\sum \limits _{n=0}^{\infty } \psi _{i}(z, t; n)\) and

$$ {\psi_{i}^{i}}(z, t)=\sum\limits_{n=0}^{\infty} \psi_{i}(z, t; 2n),\qquad \psi_{i}^{1-i}(z, t)=\sum\limits_{n=0}^{\infty} \psi_{i}(z, t; 2n+1), \quad i\in\{0, 1\}. $$

The explicit expressions for ψ0(z,t; n) and ψ1(z,t; n) can be written separately for even and odd n,n ≥ 0. Due to López and Ratanov (2014, Theorem 2.1), we have

$$ \begin{array}{@{}rcl@{}} \psi_{0}(z, t; 2n) &=\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!} t^{2n}{\varPhi}\left( n, 2n+1; 2(\upbeta-az)t\right)\exp(-({\lambda}_{0}-az)t), \end{array} $$
(A.14)
$$ \begin{array}{@{}rcl@{}} \psi_{1}(z, t; 2n) &=\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!} t^{2n}{\varPhi}(n, 2n+1; 2(az-\upbeta)t)\exp(-({\lambda}_{1}+az)t), \end{array} $$
(A.15)
$$ \begin{array}{@{}rcl@{}} \psi_{0}(z, t; 2n+1) &=\frac{{\lambda}_{0}^{n+1}{{\lambda}_{1}^{n}}}{(2n+1)!} t^{2n+1}{\varPhi}\left( n+1, 2n + 2; 2(\upbeta - az)t\right)\exp(-({\lambda}_{0} - az)t), \end{array} $$
(A.16)
$$ \begin{array}{@{}rcl@{}} \psi_{1}(z, t; 2n+1) \!&=\frac{{{\lambda}_{0}^{n}}{\lambda}_{1}^{n+1}}{(2n+1)!} t^{2n+1} {\varPhi}(n + 1, 2n+2; 2(az - \upbeta)t)\exp(-({\lambda}_{1} + az)t). \end{array} $$
(A.17)

Formulae (A.14)–(A.17) directly give the desired result (A.2)–(A.8). For instance, by differentiating in Eq. A.14 we have

$$ {\mathbb{E}_{0}^{0}}[\mathbb{T}(t)]=\sum\limits_{n=0}^{\infty}\frac{{\partial}\psi_{0}(z, t; 2n)}{{\partial} z}|_{z=0} $$
$$ =a\mathrm{e}^{-{\lambda}_{0}t}\sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!}t^{2n+1} \Big[ -2{\varPhi}^{\prime}(n, 2n+1; 2\upbeta t)+{\varPhi}(n, 2n+1; 2\upbeta t) \Big] $$

and

$$ {\mathbb{E}_{0}^{0}}[\mathbb{T}(t)^{2}]=\sum\limits_{n=0}^{\infty}\frac{{\partial}^{2}\psi_{0}(z, t; 2n)}{{\partial} z^{2}}|_{z=0} $$
$$ =a^{2}\mathrm{e}^{-{\lambda}_{0}t}\sum\limits_{n=0}^{\infty}\frac{{{\lambda}_{0}^{n}}{{\lambda}_{1}^{n}}}{(2n)!}t^{2n+2} \Big[ 4{\varPhi}^{\prime\prime}(n, 2n+1; 2\upbeta t)-4{\varPhi}^{\prime}(n, 2n+1; 2\upbeta t)+{\varPhi}(n, 2n+1; 2\upbeta t) \Big]. $$

The following known identities,

$$ {\varPhi}^{\prime}(\alpha, \upbeta; z)= \frac{\mathrm{d}{\varPhi}}{d z}(\alpha, \upbeta; z)=\frac{\alpha}{\upbeta}{\varPhi}(\alpha+1, \upbeta+1;z)$$

and

$${\varPhi}^{\prime\prime}(\alpha, \upbeta; z)=\frac{\alpha(\alpha+1)}{\upbeta(\upbeta+1)}{\varPhi}(\alpha+2, \upbeta+2;z),$$

see Gradshteyn and Ryzhik (1994, [formula 9.213]), give the result, Eqs. A.2A.10 and A.6A.12. Formulae A.3 and A.7 can be obtained similarly from Eq. A.16.

The remaining formulae of the theorem can be derived from Eqs. A.2A.3 and A.6A.7 by symmetry: Eq. A.5 follows from Eq. A.2; Eq. A.4 follows from Eq. A.3; Eq. A.9 follows from Eq. A.6; Eq. A.8 follows from Eq. A.7 after replacements \(a\rightarrow -a\) and \({\lambda }_{0}\leftrightarrow {\lambda }_{1}.\)

Equations A.2A.9 permit us to evaluate the covariance between \(\mathbb {T}(t)\) and \(\mathbb {T}(s).\)

Theorem A.3

For t > s > 0

$$ \begin{array}{@{}rcl@{}} \mathbb{E}_{0}\mathbb{T}(t)\mathbb{T}(s)=& \mathbb{E}_{0}\mathbb{T}(t-s){\cdot\mathbb{E}_{0}^{0}}\mathbb{T}(s)+\mathbb{E}_{1}\mathbb{T}(t-s){\cdot\mathbb{E}_{0}^{1}}\mathbb{T}(s)+\mathbb{E}_{0}[\mathbb{T}(s)^{2}], \end{array} $$
(A.18)
$$ \begin{array}{@{}rcl@{}} \mathbb{E}_{1}\mathbb{T}(t)\mathbb{T}(s)=& \mathbb{E}_{0}\mathbb{T}(t-s){\cdot\mathbb{E}_{1}^{0}}\mathbb{T}(s)+\mathbb{E}_{1}\mathbb{T}(t-s){\cdot\mathbb{E}_{1}^{1}}\mathbb{T}(s)+\mathbb{E}_{1}[\mathbb{T}(s)^{2}], \end{array} $$
(A.19)

where \({\mathbb {E}_{i}^{j}}[\mathbb {T}(s)], \mathbb {E}_{i}[\mathbb {T}(t-s)]\) and \(\mathbb {E}_{i}[\mathbb {T}(s)^{2}], i, j\in \{0, 1\},\) are given by Eqs. A.2A.9.

Proof

Notice that \( \mathbb {E}_{0}\mathbb {T}(t)\mathbb {T}(s)=\mathbb {E}_{0}\left [(\mathbb {T}(t)-\mathbb {T}(s))\cdot \mathbb {T}(s)\right ]+\mathbb {E}_{0}[\mathbb {T}(s)^{2}]. \)

Due to persistence and time-homogeneity of the process \(\mathbb {T},\) see Eq. 1.1,

$$ \begin{aligned} &\mathbb{E}_{0}\left[(\mathbb{T}(t)-\mathbb{T}(s))\cdot\mathbb{T}(s)\right]= \\ &\mathbb{E}[(\mathbb{T}(t)-\mathbb{T}(s)~|~{\varepsilon}(s)=0]{\cdot\mathbb{E}_{0}^{0}}\mathbb{T}(s) +\mathbb{E}[(\mathbb{T}(t)-\mathbb{T}(s)~|~{\varepsilon}(s)=1]{\cdot\mathbb{E}_{0}^{1}}\mathbb{T}(s) \\ &=\mathbb{E}_{0}\mathbb{T}(t-s){\cdot\mathbb{E}_{0}^{0}}\mathbb{T}(s)+\mathbb{E}_{1}\mathbb{T}(t-s){\cdot\mathbb{E}_{0}^{1}}\mathbb{T}(s), \end{aligned} $$

which gives Eq. A.18. Equation A.19 follows similarly. □

Remark 4

A.2 In the symmetric case λ0 = λ1 = λ > 0, the results of Theorem A.2 (Eqs. A.2A.9) and Eqs. A.18A.19 look much simpler, cf Kolesnik (2012).

Since β = 0 and Φ(⋅,⋅; 0) = 1, we have

$$ G_{n}^{(1)}(t)|_{\upbeta=0}=G_{n}^{(2)}(t)|_{\upbeta=0}\equiv\frac{1}{2n+1},\qquad H_{n}^{(1)}(t)|_{\upbeta=0}\equiv0,\qquad H_{n}^{(2)}(t)|_{\upbeta=0}\equiv\frac{1}{2n+3}. $$

Therefore, for the symmetric case, the first moments Eqs. A.2A.4 are given by

$$ \begin{aligned} {\mathbb{E}_{0}^{0}}\mathbb{T}(t)&=- {\mathbb{E}_{1}^{1}}\mathbb{T}(t)=at\mathrm{e}^{-{\lambda} t}\sum\limits_{n=0}^{\infty}\frac{({\lambda} t)^{2n}}{(2n+1)!} =at\mathrm{e}^{-{\lambda} t}\frac{\sinh{{\lambda} t}}{{\lambda} t} =\frac{a}{2{\lambda}}\left( 1-\mathrm{e}^{-2{\lambda} t}\right), \\ {\mathbb{E}_{0}^{1}}\mathbb{T}(t)&={\mathbb{E}_{1}^{0}}\mathbb{T}(t)=0, \end{aligned} $$
(A.20)

and \( \mathbb {E}_{0}\mathbb {T}(t)=-\mathbb {E}_{1}\mathbb {T}(t)=\frac {a}{2{\lambda }}\left (1-\mathrm {e}^{-2{\lambda } t}\right ). \)

The second moments are given by

$$ \begin{aligned} {\mathbb{E}_{0}^{0}}\mathbb{T}(t)^{2} &={\mathbb{E}_{1}^{1}}\mathbb{T}(t)^{2}\\ &=(at)^{2}\mathrm{e}^{-{\lambda} t}\sum\limits_{n=0}^{\infty}\frac{({\lambda} t)^{2n}}{(2n+1)!}=(at)^{2}\mathrm{e}^{-{\lambda} t}\frac{\sinh{{\lambda} t}}{{\lambda} t} =\frac{a^{2}t}{2{\lambda}} \left( 1-\mathrm{e}^{-2{\lambda} t}\right), \\ {\mathbb{E}_{0}^{1}}\mathbb{T}(t)^{2} &={\mathbb{E}_{1}^{0}}\mathbb{T}(t)^{2}\\ &=(at)^{2}\mathrm{e}^{-{\lambda} t}\sum\limits_{n=0}^{\infty}\frac{({\lambda} t)^{2n+1}}{(2n+1)!(2n+3)} =(at)^{2}\mathrm{e}^{-{\lambda} t}\left( \frac{\sinh z-z}{z}\right)'|_{z={\lambda} t}\\ &=\frac{a^{2}t}{2{\lambda}}\left( 1+\mathrm{e}^{-2{\lambda} t}\right)-\frac{a^{2}}{2{\lambda}^{2}}\left( 1-\mathrm{e}^{-2{\lambda} t}\right), \end{aligned} $$

and, by summing we have

$$ \mathbb{E}_{0}\mathbb{T}(t)^{2}= {\mathbb{E}_{0}^{0}}\mathbb{T}(t)^{2}+ {\mathbb{E}_{0}^{1}}\mathbb{T}(t)^{2} =\frac{a^{2}}{2{\lambda}^{2}}\left( \mathrm{e}^{-2{\lambda} t}-1+2{\lambda} t\right)=\mathbb{E}_{1}\mathbb{T}(t)^{2}. $$
(A.21)

Equations A.20 and A.21 are consistent with known results, see e.g. Kolesnik and Ratanov (2013, (4.2.24)).

By Eqs. A.18A.19A.20 and A.21,

$$\begin{aligned} &\mathbb{E}_{0}[\mathbb{T}(t)\mathbb{T}(s)]=\mathbb{E}_{1}[\mathbb{T}(t)\mathbb{T}(s)]\\ &=\frac{a}{2{\lambda}}\left( 1-\mathrm{e}^{-2{\lambda}(t-s)}\right)\cdot\frac{a}{2{\lambda}}\left( 1-\mathrm{e}^{-2{\lambda} s}\right) +0+\frac{a^{2}}{2{\lambda}^{2}}\left( \mathrm{e}^{-2{\lambda} s}-1+2{\lambda} s\right) \end{aligned}$$
$$ =\frac{a^{2}}{4{\lambda}^{2}}\left[4{\lambda} s-(1+\mathrm{e}^{-2{\lambda}(t-s)})(1-\mathrm{e}^{-2{\lambda} s})\right], $$
(A.22)

and the covariance becomes

$$\begin{aligned} \text{cov}(\mathbb{T}(t), \mathbb{T}(s))=&\mathbb{E}[\mathbb{T}(t)\cdot\mathbb{T}(s)]-\mathbb{E}[\mathbb{T}(t)]\cdot\mathbb{E}[\mathbb{T}(s)]\\ =&\frac{a^{2}}{2{\lambda}^{2}} \left[2{\lambda} s-1+\mathrm{e}^{-2{\lambda} s}+\mathrm{e}^{-2{\lambda} t}-\frac12\left( \mathrm{e}^{-2{\lambda}(t-s)}+\mathrm{e}^{-2{\lambda}(t+s)}\right)\right]. \end{aligned}$$

It is known that under Kac’s scaling, \(a, {\lambda }\to \infty , a^{2}/{\lambda }\to \sigma ^{2},\) see Kac (1974), Kolesnik and Ratanov (2013), and López and Ratanov (2012), the symmetric telegraph process \(\mathbb {T}(t)\) converges to Brownian motion σW(t). Equations A.20A.21 and A.22 consist with this convergence: under this scaling we have

  • by Eq. A.20,

    $$ \lim\mathbb{E}_{0}[\mathbb{T}(t)]=\lim\mathbb{E}_{1}[\mathbb{T}(t)]=0; $$
  • by Eq. A.21,

    $$ \lim\mathbb{E}_{0}[\mathbb{T}(t)^{2}]=\lim\mathbb{E}_{1}[\mathbb{T}(t)^{2}]=\sigma^{2}t, $$
  • by Eq. A.22,

    $$ \lim\mathbb{E}_{0}[\mathbb{T}(t)\mathbb{T}(s)]=\lim\mathbb{E}_{1}[\mathbb{T}(t)\mathbb{T}(s)]=\sigma^{2}s,\qquad s\leq t. $$

Remark 2

Notice that the “general” telegraph process \(\mathbb {T}(t), t\geq 0,\) with two alternating velocities a0 and a1,a0 > a1, can be reduced to the symmetric case:

$$ \mathbb{T}(t)\stackrel{D}{=}(a_{0}+a_{1})t/2+\mathbb{T}_{\pm a}^{\text{sym}}(t), $$

where \(\mathbb {T}_{\pm a}^{\mathrm sym}(t)\) is the telegraph process with symmetric velocities ± a,a = (a0a1)/2. Therefore, without loss of generality, only a “symmetric” process \(\mathbb {T}(t), t\geq 0,\) can be studied.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ratanov, N. Ornstein-Uhlenbeck Processes of Bounded Variation. Methodol Comput Appl Probab 23, 925–946 (2021). https://doi.org/10.1007/s11009-020-09794-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11009-020-09794-x

Keywords

Mathematics Subject Classification (2010)

Navigation