Publication Cover
Stochastics
An International Journal of Probability and Stochastic Processes
Volume 95, 2023 - Issue 1
1,060
Views
2
CrossRef citations to date
0
Altmetric
Research Article

Standard and fractional reflected Ornstein–Uhlenbeck processes as the limits of square roots of Cox–Ingersoll–Ross processes

ORCID Icon &
Pages 99-117 | Received 24 Sep 2021, Accepted 22 Feb 2022, Published online: 17 Mar 2022

Abstract

In this paper, we establish a new connection between Cox–Ingersoll–Ross (CIR) and reflected Ornstein–Uhlenbeck (ROU) models driven by either a standard Wiener process or a fractional Brownian motion with H>12. We prove that, with probability 1, the square root of the CIR process converges uniformly on compacts to the ROU process as the mean reversion parameter tends to either σ2/4 (in the standard case) or to 0 (in the fractional case). This also allows to obtain a new representation of the reflection function of the ROU as the limit of integral functionals of the CIR processes. The results of the paper are illustrated by simulations.

MSC 2020:

1. Introduction

Both the reflected Ornstein–Uhlenbeck (ROU) and the Cox–Ingersoll–Ross (CIR) processes are extremely popular models in a variety of fields. Without attempting to give a complete overview of possible applications due to a large amount of literature on the topic, we only mention that the ROU process is widely used in queueing theory [Citation10,Citation30–32], in population dynamics modelling [Citation1,Citation25], in economics and finance for modelling regulated markets [Citation3,Citation4,Citation17,Citation33], interest rates [Citation11] and stochastic volatility [Citation27] (see also Refs [Citation9,Citation20] and references therein for more details on applications of the ROU in various fields) while the most notable usages of the CIR process are related to representing the dynamics of interest rates [Citation5–7] and stochastic volatility in the Heston model [Citation12].

It is well-known [Citation21,Citation28] that the CIR process has strong links with the standard OU dynamics; in particular, if B=(B1,,Bd) is a d-dimensional Brownian motion and U=(U1,,Ud) is a standard d-dimensional OU process given by Ui(t)=Ui(0)b20tUi(s)ds+σ2Bi(t),t0,i=1,,d,then it is easy to see via Itô's formula that the process i=1dUi2(t), t0, is the CIR process of the form (1) X(t)=X(0)+0t(abX(s))ds+σ0tX(s)dW(s),t0,(1) with a=dσ24 and W(t):=i=1d0tUi(s)j=1dUj2(s)dBi(s) (which is a standard Brownian motion by Levy's characterization). The value d=4aσ2 is sometimes referred to as a dimension or a number of degrees of freedom of the CIR process (see e.g. Ref. [Citation21] and references therein) and thus, in this terminology, a square of a standard one-dimensional OU process turns out to be a CIR process with one degree of freedom w.r.t. another Brownian motion.

In this paper, we investigate a connection between the CIR and the ROU processes that is in some sense related to the one described above. Namely, in the first part we prove that the ROU process (2) Y(t)=Y(0)b20tY(s)ds+σ2W(t)+L(t),t0,(2) where W is a standard Brownian motion and L is a continuous non-decreasing process that can have points of growth only at zeros of Y, coincides with the square root of the CIR process of the type (Equation1) with a=σ24 (i.e. with one degree of freedom) driven by the same Brownian motion W. Moreover, if {εn,n1} is a sequence of positive numbers such that εn0 as n, then, with probability 1, for all T>0 (3) supt[0,T]|L(t)120tεnXεn(s)ds|0,n,(3) where Xεn is the CIR process of the form Xεn(t)=X(0)+0t(σ24+εnbXεn(s))ds+σ0tXεn(s)dW(s).

The second part of the paper discusses the connection between fractional counterparts of Equations (Equation1) and (Equation2) driven by fractional Brownian motion {BH(t),t0} with Hurst index H>12. Namely, we consider a fractional Cox–Ingersoll–Ross process XεH(t)=X(0)+0t(εbXεH(s))ds+σ0tXεH(s)dBH(s),t0,where the integral 0tXH(s)dBH(s) is understood as the pathwise limit of Riemann-Stieltjes integral sums (see [Citation23] or [Citation8, Subsection 4.1]) and prove that with probability 1 the paths of {XεH(t),t0} a.s. converge to the reflected fractional Ornstein–Uhlenbeck (RFOU) process uniformly on each compact [0,T] as ε0. Moreover, an analogue of the representation (Equation3) also takes place: if LH is a reflection function of the RFOU process, then, with probability 1, for each T>0 supt[0,T]|LH(t)120tεXεH(s)ds|0,ε0.

The paper is organised as follows. In Section 2, we consider the link between the CIR and the ROU processes in the standard Wiener case. Section 3 is devoted to the fractional setting. Section 4 contains simulations that illustrate our results.

2. Classical reflected Ornstein–Uhlenbeck and Cox–Ingersoll–Ross processes

The main goal of this section is to establish a connection between Cox–Ingersoll–Ross (CIR) and reflected Ornstein–Uhlenbeck (ROU) processes in the standard Brownian setting. We shall start from the definition of a reflection function following the one given in the classical work [Citation29].

Definition 2.1

Let ξ={ξ(t),t0} be some stochastic process. The process ζ={ζ(t),t0} is called a reflection function for ξ, if ζ is, with probability 1, a continuous non-decreasing process such that ζ(0)=0 and the points of growth of ζ can occur only at zeros of ξ.

Definition 2.2

Stochastic process Y~={Y~(t),t0} is called a reflected Ornstein–Uhlenbeck (ROU) process if it satisfies a stochastic differential equation of the form (4) Y~(t)=Y(0)b~0tY~(s)ds+σ~W(t)+L~(t),t0,(4) where Y(0), b~ and σ~ are positive constants, W={W(t),t0} is a standard Brownian motion, {L~(t),t0} is a reflection function for Y~ and Y~0 a.s.

Remark 2.1

The ROU process is well-known and studied in the literature, see e.g. [Citation31] and references therein. Note also that, despite (Equation4) has two unknown functions Y~ and L~, the solution is still unique. Indeed, let Y~ and Y^ be two stochastic processes satisfying Y~(t)=Y(0)b~0tY~(s)ds+σ~W(t)+L~(t)and Y^(t)=Y(0)b~0tY^(s)ds+σ~W(t)+L^(t),where L~ and L^ are the corresponding reflection functions. Assume that on some ωΩ such that both Y~ and Y^ are continuous (5) Y~(t)Y^(t)>0(5) and consider τ(t):=sup{s[0,t):Y~(t)Y^(t)=0}. Then Y~(u)Y^(u)>0 for all u(τ(t),t]; moreover, Y~(u)>0 for u(τ(t),t], therefore L~ is non-increasing on (τ(t),t]. It means that the difference Y~(u)Y^(u) is also non-increasing on (τ(t),t] since Y~(u)Y^(u)=b~τ(t)u(Y~(s)Y^(s))ds+(L~(u)L^(u))(L~(τ(t))L^(τ(t)))and the right-hand side is non-increasing w.r.t. u. Whence, taking into account that Y~(τ(t))Y^(τ(t))=0 due to the definition of τ(t) and continuity of both Y~ and Y^, the difference Y~(u)Y^(u) cannot be positive for any u(τ(t),t] which contradicts (Equation5). Interchanging the roles of Y~ and Y^, one can easily verify that Y~(t)Y^(t) cannot be negative either and whence Y^=Y~, L^=L~.

Now, consider a standard CIR process defined as a continuous modification of the unique solution to the equation (6) X(t)=X(0)+0t(abX(s))ds+σ0tX(s)dW(s),t0,(6) where X(0),a,b,σ>0 and W={W(t),t0} is a classical Wiener process. It is well-known (see e.g. [Citation15, Example 8.2]) that for a>0 the solution {X(t),t0} is non-negative a.s. for any t0; moreover, the solution is strictly positive a.s. provided that aσ22, see e.g. [Citation16, Chapter 5]. Therefore, if a>0, the square-root process Y={Y(t),t[0,T]}:={X(t),t[0,T]} is well-defined.

For an arbitrary ε>0, consider a stochastic process {X(t)+ε,t[0,T]}. By Itô's formula, for any t0 (7) X(t)+ε=X(0)+ε+120t(aX(s)+εσ24X(s)(X(s)+ε)32)ds120tbX(s)X(s)+εds+σ20tX(s)X(s)+εdW(s)(7) and, since the left-hand side of (Equation7) converges to X(t)=Y(t) a.s. as ε0, moving ε0 in the right-hand side would give us the dynamics of Y.

First, it is clear that for any t0 (8) X(0)+εY(0)a.s.(8) and (9) 0tX(s)X(s)+εds0tY(s)dsa.s.(9) as ε0. Further, by the monotone convergence, (10) 0t1X(s)+εds0t1Y(s)ds[0,){}a.s.,0tX(s)(X(s)+ε)320t1Y(s)ds[0,){}a.s.(10) as ε0. Finally, by Burkholder–Davis–Gundy inequality and dominated convergence theorem, for any T>0 E(supt[0,T]|0tX(s)X(s)+εdW(s)W(t)|)24E0T(X(s)X(s)+ε1)2ds=4E0T(X(s)X(s)+ε1)21{X(s)>0}ds+4E0T1{X(s)=0}ds=4E0T(X(s)X(s)+ε1)21{X(s)>0}ds0,ε0,where we used continuity of the distribution of X(s) for each s>0 to state that 4E0T1{X(s)=0}ds=0 (see e.g. [Citation21] and references therein). This implies that (11) supt[0,T]|0tX(s)X(s)+εdW(s)W(t)|L2(Ω)0,ε0.(11)

By (Equation11), it is evident that there exists a sequence {εn,n1} which depends on T such that (12) supt[0,T]|0tX(s)X(s)+εndW(s)W(t)|0a.s.,n(12) and along this sequence (13) limn120t(aX(s)+εnσ24X(s)(X(s)+εn)32)ds<,t[0,T],(13) a.s. because all other limits in (Equation7) as εn0 are finite a.s. However, the integral 0t1Y(s)ds that arises in (Equation10) may be infinite and thus the explicit form of the limit above, for now, remains obscure. This issue as well as the connection of Y to the ROU process is addressed in the next theorem.

Theorem 2.1

Let Y={Y(t),t0}={X(t),t0} be the square root process, where X is the CIR process defined by (Equation6). τ:=inf{t0:X(t)=0}=inf{t0:Y(t)=0}.

  1. If a>σ24, then for any t0 0t1Y(s)ds<a.s.

    Moreover, the square root process Y a.s. satisfies the SDE of the form (14) Y(t)=Y(0)+12(aσ24)0t1Y(s)dsb20tY(s)ds+σ2W(t),(14)

    Y(0)=X(0), and the solution to this equation is unique among non-negative stochastic processes.

  2. If a=σ24, then 0τ1Y(s)ds<a.s.while 0τ+γ1Y(s)ds=a.s.for any γ>0. Moreover, the square root process Y satisfies the SDE of the form (15) Y(t)=Y(0)b20tY(s)ds+σ2W(t)+L(t),(15) where the process L from (Equation15) is a continuous non-decreasing process the points of growth of which can occur only at zeros of Y, i.e. Y is a reflected Ornstein–Uhlenbeck process.

Proof.

Case (a): a>σ24. Denote p:=aσ24>0. Our goal is to prove that the integral 0t1X(s)ds=0t1Y(s)dsis finite a.s. Define A(t):={ωΩ:0t1X(s)ds=+} and assume that for some t>0: P(A(t))>0. Fix T>t, the corresponding sequence {εn,n1} such that convergence (Equation12) holds and an arbitrary ωA(t)Ω, where ΩΩ, P(Ω)=1, is the set where (Equation12) takes place (in what follows, ω in brackets will be omitted). Then 0t(aX(s)+εnσ24X(s)(X(s)+εn)32)ds=σ240t(1X(s)+εnX(s)(X(s)+εn)32)ds+p0t1X(s)+εnds.

Obviously, for all s[0,t] 1X(s)+εnX(s)(X(s)+εn)32a.s.,so, for ωA(t)Ω 0t(aX(s)+εnσ24X(s)(X(s)+εn)32)dsa.s.,n,whence, taking into account (Equation7)–(Equation9), we obtain that X(t)X(0)+b20tX(s)dsσ2W(t)=,which is impossible a.s. We get a contradiction, whence P(A(t))=0 for all t0 and 0t1X(s)ds=0t1Y(s)ds< a.s. By going to the limit in (Equation7), we immediately get (Equation14).

Concerning the uniqueness of solution to (Equation14), let Y~(t) be any of its non-negative solutions. Then, by Itô's formula, Y~2(t)=X(0)+0t(abY~2(s))ds+σ0tY~(s)dW(s)so Y~ satisfies Equation (Equation6) and thus coincides with X. Therefore Y~(t)=X(t)=Y(t)a.s.,t0.Case (b): a=σ24. Fix T>0 and take the corresponding sequence {εn,n1} such that (Equation12) holds. By (Equation7), for any t[0,T] X(t)+εn=X(0)+εn+120t(aX(s)+εnσ24X(s)(X(s)+εn)32)ds120tbX(s)X(s)+εnds+σ20tX(s)X(s)+εndW(s)=X(0)+εn+σ280tεn(X(s)+εn)32ds120tbX(s)X(s)+εnds+σ20tX(s)X(s)+εndW(s),and (Equation13) implies that there exists ΩΩ, P(Ω)=1, such that for all ωΩ the limit L(t):=limnσ280tεn(X(s)+εn)32dsis well-defined and finite for all t[0,T]. It is evident that L(0)=0 a.s. due to continuity of X and the fact that X(0)>0. Moreover, since a.s. L(t)=X(t)X(0)+b20tX(s)dsσ2W(t)t[0,T],

L is continuous in t. Furthermore, 0t1εn(X(s)+εn)32ds0t2εn(X(s)+εn)32dsfor all t1<t2, and whence L is non-decreasing in t a.s. Finally, if X(t)=x>0, there exists an interval [t1,t2] containing t such that X(s)>x2 for all s[t1,t2] and thus L(t1)L(t2)=limnσ28t1t2εn(X(s)+εn)32ds0,n,i.e. L can increase only at points of zero hitting of X that coincide with the ones of Y. Taking into the account all of the above as well as an arbitrary choice of T, L is the reflection function for Y and the latter is indeed a ROU process.

Now, let us prove that 0τ1Ysds< a.s. Consider a standard Ornstein–Uhlenbeck process U={U(t),t0} of the form (16) U(t)=X(0)b20tU(s)ds+σ2W(t),(16) with W being the same Brownian motion that drives X. It is evident that Y coincides with U until τ a.s. and thus it is sufficient to prove that 0τ1U(s)ds< a.s. For any ε>0 consider σ240τ1U(s)1{ε<U(s)<1}ds=ε1LU(τ,x)xdx,where LU denotes the local time of U, and observe that σ240τ1U(s)dslimε00τ1U(s)1{ε<U(s)<1}ds=01LU(τ,x)xdx.

Computations similar to the ones in [Citation26, Section IV.44] indicate that the local time LU(t,x) of U is Hölder continuous in x up to order 12 over bounded time intervals and thus 0τ1U(s)ds=0τ1Y(s)ds< a.s.

Finally, assume that for some γ>0, 0τ+γ1Y(s)ds<with positive probability. On ωΩ where this property holds, we have that 0τ+γdsX(s)+ε0τ+γX(s)(X(s)+ε)32ds0τ+γ1Y(s)ds0τ+γ1Y(s)ds=0,ε0.

Therefore, for such ω, Y satisfies the equation of the form Y(t)=Y(0)b20tY(s)ds+σ2W(t)on the interval [0,τ+γ], i.e. such paths of Y coincide with the corresponding paths of the Ornstein–Uhlenbeck process U defined by (Equation16) up until τ+γ. This implies that U(τ)=0 and U is non-negative on the interval [τ,τ+γ] for such ω, which is impossible due to the non-tangent property of Gaussian processes stated by [Citation34], see also [Citation24].

Remark 2.2

Since the integral 0t1X(s)ds is finite a.s. for a>σ24, X(t)+εX(0)+ε120t(aX(s)+εσ24X(s)(X(s)+ε)3/2)ds+120tbX(s)X(s)+εdsa.s.X(t)X(0)12(aσ24)0t1X(s)ds+b20tX(s)ds<as ε0. Therefore, taking into account (Equation7), 0tX(s)X(s)+εdW(s)a.s.W(t),ε0.

As a corollary of Theorem 2.1, we have a representation of the reflection function of the ROU process as the limit of integral functionals of the CIR processes. It is interesting that the reflection function is singular w.r.t. the Lebesgue measure (see Remark 2.3) while the processes that converge to it are absolutely continuous a.s.

Theorem 2.2

Let {W(t),t0}} be a continuous modification of a standard Brownian motion, Y(0), b, σ>0 be given constants and {εn,n1} be an arbitrary sequence such that εn0, n. For any εn from this sequence, consider the CIR process Xεn={Xεn(t),t0} given by Xεn(t)=X(0)+0t(σ24+εnbXεn(s))ds+σ0tXεn(s)dW(s)and denote its square root by Yεn(t):=Xεn(t). Then, with probability 1,

  1. the limit limnYεn(t)=:Y(t) is well-defined, finite and non-negative for any t0;

  2. the limit process Y={Y(t),t0} is a ROU process satisfying the equation of the form Y(t)=Y(0)b20tY(s)ds+σ2W(t)+L(t),t0,with Y(0)=X(0)>0 and L being the reflection function for Y;

  3. for any T>0 (17) supt[0,T]|Y(t)Yεn(t)|0,n,(17) and (18) supt[0,T]|L(t)120tεnYεn(s)ds|0,n.(18)

Proof.

Denote by X the CIR process of the form X(t)=X(0)+0t(σ24bX(s))ds+σ0tX(s)dW(s).

By Theorem 2.1, there exists ΩΩ, P(Ω)=1, such that for all ωΩ X and each Yεn, n1, are continuous and the latter satisfy equations of the form Yεn(t)=Y(0)+120tεnYεn(s)dsb20tY(s)ds+σ2W(t),t0,with the integral 0t1Yεn(s)ds<. Furthermore, since each Xεn=Yεn2 is a CIR process that satisfies conditions of the comparison theorem from [Citation14], this Ω can be chosen such that for all ωΩ (19) Yεn(ω,t)Yεn+1(ω,t)X(t)0,t0,n1.(19)

Fix ωΩ (in what follows, we will omit ω in brackets for notational simplicity). Since the sequence {Yεn(t),n1} is non-increasing for each t0, there exists a pointwise limit Y(t):=limnYεn(t)[0,). Moreover, it is evident that limn0tYεn(s)ds=0tY(s)ds and since Y(t)=limnYεn(t)=Y(0)limnb20tYεn(s)ds+σ2W(t)+limn120tεnYεn(s)ds=Y(0)b20tY(s)ds+σ2W(t)+limn120tεnYεn(s)ds,the limit L(t):=limn120tεnYεn(s)ds is well-defined, non-negative and finite.

In order to obtain the claim of the theorem, it is sufficient to check that the function L defined above is indeed a reflection function for Y, i.e. is continuous and non-decreasing process that starts at zero and the points of growth of which occur only at zeros of Y. Note that continuity of L would also imply the uniform convergences (Equation17) and (Equation18) on each compact [0,T]. Indeed, since Yεn(t)Yεn+1(t) for all t0, n1 and continuity of L would imply continuity of Y, Dini's theorem guarantees (Equation17). The same argument applies to (Equation18): the right-hand side of 120tεnYεn(s)ds=Yεn(t)Y(0)+b20tYεn(s)dsσ2W(t)is non-increasing w.r.t. t, therefore for each t0 and n1 120tεnYεn(s)ds120tεn+1Yεn+1(s)dsand Dini's theorem implies (Equation18) as well.

By (Equation19), continuity of X and the fact that X(0)>0, there exists an interval [0,t0) such that for all t[0,t0) and n1 Yεn(t)Y(0)2. Thus for any t[0,t0) L(t)=limn120tεnYεn(s)dslimnt0εnY(0)=0,i.e. L(t)=0 for all t[0,t0].

For the reader's convenience, we will split the further proof into four steps.

Step 1: L in non-decreasing. Monotonicity of L is obvious since for any fixed n1 and t1<t2 0t1εnYεn(s)ds0t2εnYεn(s)ds.

Step 2: right-continuity. Let us show that L is continuous from the right. For any fixed t0, denote L(t+):=limδ0L(t+δ) (the right limit exists since L is non-decreasing) and assume that L(t+)L(t)=α>0. Due to the monotonicity of L, this implies that for all δ>0 (20) L(t+δ)L(t)α>0.(20)

Now, take n0 such that for all nn0 120tεnYεn(s)ds[L(t),L(t)+α4)and δ0>0 such that 120t+δ0εn0Yεn0(s)ds[L(t),L(t)+α2).

As it was noted previously, for each s0 the values of 120sεnYεn(u)du are non-increasing when n. Thus for any nn0 120t+δ0εnYεn(s)ds120t+δ0εn0Yεn0(s)dsL(t)+α2,i.e. L(t+)limn120t+δ0εnYεn(s)ds<L(t)+α2,which contradicts (Equation20). Therefore, L(t+)L(t)=0, i.e. L is right-continuous.

Step 3: left-continuity. Now, let us show that L is continuous from the left. Assume that it is not true and there exists t>0 such that L(t)L(t)>0 (note that L(t)=limδ0L(tδ) is well-defined due to the monotonicity of L). Since L may have only positive jumps, so does Y and, moreover, the points of jumps of L and Y coincide. This implies that Y(t)Y(t)>0 and we now consider two cases.

Case 1: Y(t)=y>0. Then Y(t)=Y(t+)>y (note that Y is right-continuous by Step 2) and there exists an interval [tδ,t+δ] such that Yεn(s)Y(s)>y2 for all s[tδ,t+δ]. This implies that L(t+δ)L(tδ)=limn12tδt+δεnYεn(s)dslimn2δεny=0,i.e. L cannot have a jump at t. This means that Y cannot have a jump at point t either and we obtain a contradiction.

Case 2: Y(t)=0 and Y(t+)=Y(t)=y>0. Fix T>t, λ(0,12) and let Λ be a random variable such that for all t1,t2[0,T] |W(t1)W(t2)|Λ|t1t2|λ.

Take n11 and δ1>0 such that εn1+δ1+σΛ2δ1λ<y and note that there exists δ2<δ1 such that Y(tδ2)<εn1. Since Yεn(tδ2)Y(tδ2) as n, there exists n2>n1 such that Yεn2(tδ2)<εn1<y. Moreover, Yεn2(t)Y(t)=y thus one can define τ:=sup{s(tδ2,t),Yεn2(s)=εn1}.

Observe that Yεn2(τ)=εn1 and Yεn2(s)εn1 for all s[τ,t], whence Yεn2(t)=Yεn2(τ)+12τtεn2Yεn2(s)dsb2τtYεn2(s)ds+σ2(W(t)W(τ))εn1+εn22εn1(tτ)+σΛ2(tτ)λεn1+δ1+σΛ2δ1λ<y,which contradicts the assumption that Yεn2(t)y. This contradiction together with all of the above implies that Y (and thus L) is continuous at each point t0.

Step 4: points of growth. Now, let us prove that the points of growth of L may occur only at zeros of Y. Indeed, let t>0 be such that Y(t)=y>0. Since Y is continuous, there exists δ3>0 such that for any s(tδ3,t+δ3) Y(s)>y2>0.

This, in turn, implies that for all s(tδ3,t+δ3) and n1 Yεn(s)Y(s)>y2>0and thus for any δ[0,δ3) L(t+δ)L(tδ)=limn12tδt+δεnYεn(s)dslimn2δyεn=0.

Therefore L(t+δ)L(tδ)=0 and L does not grow in some neighbourhood of t.

Remark 2.3

It is well-known (see e.g. [Citation2, Appendix A] or [Citation37, Subsection 3.3.1]) that the absolute value of OU and ROU processes with non-zero mean reversion levels do not coincide. In turn, in the ‘symmetric’ case with zero mean reversion parameter, the absolute value of the OU process and ROU process have the same distribution but do not coincide pathwisely. Theorem 2.1 allows to clarify this subtle difference in the following manner.

Let B={B(t),t0} be some standard Brownian motion and U(t)=U(0)b20tU(s)ds+σ2B(t),t0,be a standard Ornstein–Uhlenbeck process with non-random positive initial value U(0)>0. By Itô's formula, U2(t)=U2(0)+0t(σ24bU2(s))ds+σ0tU(s)dB(s)=U2(0)+0t(σ24bU2(s))ds+σ0t|U(s)|sign(U(s))dB(s)=U2(0)+0t(σ24bU2(s))ds+σ0t|U(s)|dW(s),where W(t):=0tsign(U(s))dB(s) is a standard Brownian motion (which can be easily verified by Levy's characterization). Thus, the process X(t):=U2(t), t0, is a CIR process w.r.t. W. By Theorem 2.1, the square root process Y(t):=X(t), t0, is a reflected Ornstein–Uhlenbeck process with respect to W satisfying the SDE of the form (21) Y(t)=U(0)b20tY(s)ds+σ2W(t)+L(t),t0,(21) with L being the reflection function for Y. Since Y(t)=X(t)=|U(t)|, by Tanaka's formula (22) Y(t)=U(0)+0tsign(U(s))dU(s)+LU(t)=U(0)b20tsign(U(s))U(s)ds+σ20tsign(U(s))dB(s)+LU(t)=U(0)b20tY(s)ds+σ2W(t)+LU(t),(22) with LU being the local time of U at zero. Comparing (Equation21) and (Equation22), we obtain that L(t)=LU(t), i.e. the reflection function of the ROU process Y coincides with local time at zero of the OU process U.

3. Fractional Cox–Ingersoll–Ross and fractional reflected Ornstein–Uhlenbeck processes

Let now {BH(t),t0} be a continuous modification of a fractional Brownian motion with Hurst index H>12. Consider a stochastic differential equation of the form (23) YH(t)=Y(0)+120t(aYH(s)bYH(s))ds+σ2dBH(t),t0,(23) where Y(0)>0 is a given constant, a, b, σ>0. According to [Citation23] (see also [Citation8]), the SDE (Equation23) a.s. has a unique pathwise solution {YH(t),t0} such that YH(t)>0 for all t0, and the subset of Ω where this solution exists does not depend on Y(0), a, b or σ (in fact, the solution exists for all ωΩ such that BH(ω,t) is locally Hölder continuous in t). Moreover, it can be shown (see [Citation23, Theorem 1] or [Citation8, Subsection 4.1]) that the process XH(t)=(YH(t))2, t0, satisfies the SDE of the form (24) XH(t)=X(0)+0t(abXH(s))ds+σ0tXH(s)dBH(s),t0,(24) where X(0)=Y2(0) and the integral with respect to the fractional Brownian motion exists as the pathwise limit of the corresponding Riemann-Stieltjes integral sums. Taking into account the form of (Equation24), the process {XH(t),t0} can be interpreted as a natural fractional generalization of the Cox–Ingersoll–Ross process with {YH(t),t0} being its square root.

Remark 3.1

It is evident that the solution to (Equation24) is unique in the class of non-negative stochastic processes with paths that are Hölder-continuous up to the order H. Indeed, by the fractional pathwise counterpart of the Itô's formula (see e.g. [Citation35, Theorem 4.3.1]) the square root of the solution must satisfy the equation (Equation23) until the first moment of zero hitting. However, as it was noted above, the solution to (Equation23) is unique and strictly positive a.s., i.e. never hits zero.

Now, let us recall the definition of the reflected fractional Ornstein–Uhlenbeck (RFOU) process.

Definition 3.1

Stochastic process Y~H={Y~H(t),t0} is called a fractional reflected Ornshein–Uhlenbeck (RFOU) process if it satisfies a stochastic differential equation of the form (25) Y~H(t)=Y(0)b~0tY~H(s)ds+σ~BH(t)+L~H(t),t0,(25) where Y(0), b~ and σ~ are positive constants, BH={BH(t),t0} is a fractional Brownian motion, {L~H(t),t0} is a reflection function for Y~H in the sense of Definition 2.1 and Y~H0 a.s.

Remark 3.2

For more details on properties of the RFOU process see e.g. [Citation19] and references therein. Note that, by the argument similar to the one stated in Remark 2.1, the solution (YH,LH) to the Equation (Equation25) is unique.

When it comes to the connection between FCIR and RFOU processes, there is a notable difference from the standard Brownian case discussed in section 2: in the standard case, the ROU process turned out to coincide with the square root of the CIR process with a=σ24 which is not true for the fractional case. More precisely, if a>0, XH is strictly positive a.s. and thus XH cannot coincide with the RFOU process. Furthermore, for a = 0 [Citation22, Theorem 6] claims existence and uniqueness of the solution to (Equation24) when H(23,1), and this solution turns out to stay in zero after hitting it, i.e. its square root is also different from the RFOU process. However, it is still possible to establish a clear connection between FCIR and RFOU processes highlighted in the next theorem.

Theorem 3.1

Let {BH(t),t0} be a continuous modification of a fractional Brownian motion with Hurst index H>12, Y(0), b, σ>0 be given constants. For any ε>0, consider a square root process YεH={YεH(t),t0} given by YεH(t)=Y(0)+120t(εYεH(s)bYεH(s))ds+σ2BH(t),t0.

Then, with probability 1,

  1. the limit limε0YεH(t)=:YH(t) is well-defined, finite and non-negative for any t0;

  2. the limit process YH={YH(t),t0} is a RFOU process satisfying the equation of the form (26) YH(t)=Y(0)b20tYH(s)ds+σ2BH(t)+LH(t),t0;(26)

  3. for any T>0 supt[0,T]|YH(t)YεH(t)|0,ε0,and supt[0,T]|LH(t)120tεYεH(s)ds|0,ε0.

Proof.

Let ωΩ such that BH(ω,t) is locally Hölder continuous in t be fixed (for notational simplicity, we will omit it in brackets). As it was noted above, for such ω all YεH are well-defined and strictly positive. Moreover, by the comparison theorem (see e.g. [Citation23, Lemma 1] or [Citation8, Lemma A.1]), for all t0 and ε1>ε2 Yε1H(t)>Yε2H(t)>0.

This implies that for any fixed t0 the limits limε0YεH(t)=YH(t) and limε0120tεYεH(s)ds=:LH(t) are well-defined, non-negative and finite. Furthermore, by comparison theorem, each YεH exceeds the fractional Ornstein–Uhlenbeck of the form UH(t)=Y(0)b20tUH(s)ds+σ2BH(t),t0,and hence there exists an interval [0,t0) such that for all t[0,t0) and ε>0 it holds that YεH(t)Y(0)2. Thus for any t[0,t0) LH(t)=limε0120tεYεH(s)dslimε0t0εY(0)=0,i.e. LH(t)=0 for all t[0,t0).

The remaining part of the proof is identical to the one of Theorem 2.2.

Remark 3.3

Theorem 3.1 and the preceding remark highlight that the FCIR process (Equation24) is not continuous at zero w.r.t. the mean-reversion parameter a.

4. Simulations

Let us illustrate the results with simulations. On Figure , the black paths depict simulated trajectories of the square root {YεH(t),t0} of the FCIR process given by an equation of the form YεH(t)=Y(0)+120tεYεH(s)dsb20tYεH(s)ds+σ2BH(t)with Y(0)=0.25, b=1, σ=4, ε=0.0001 and different Hurst indices H; the bold grey lines are the corresponding integrals 120tεYεH(s)ds. In order to simulate YεH, the backward Euler approximation technique from [Citation18] was used, see also [Citation13,Citation36].

Figure 1. Sample paths of YεH(t) (black) and 120tεYεH(s)ds (grey) for ε=0.0001 and different Hurst indices H. (a) H = 0.6 (b) H = 0.7 (c) H = 0.8 (d) H = 0.9

Figure 1. Sample paths of YεH(t) (black) and 12∫0tεYεH(s)ds (grey) for ε=0.0001 and different Hurst indices H. (a) H = 0.6 (b) H = 0.7 (c) H = 0.8 (d) H = 0.9

Theorem 3.1 states that the grey line approximates the reflection function LH of the RFOU process and it can be clearly seen that the plot is well agreed with the theory: the integral 120tεYεH(s)ds shows notable growth only when the corresponding path of YεH is very close to zero.

Figure  illustrates the uniform convergence of paths of YεH to the path of RFOU process as ε0. On the picture, H = 0.6, Y(0)=0.25, b = 1, σ=4 and the path of the FROU process YH was simulated using the Euler-type method: YH(0)=Y(0),YH(tn+1)=max{0,YH(tn)b2YH(tn)(tn+1tn)+σ2(BH(tn+1)BH(tn))}.

Figure 2. Comparison of the YεH with ε=1 , ε=0.5 , ε=0.25 ε=0.1 , ε=0.0001 (lines with colors ranging from light grey to dark grey) and the RFOU process (bold black line). Note that the purple path (ε=0.0001) is not visible on the plot since it almost completely coincides with the bold black trajectory of the RFOU process.

Figure 2. Comparison of the YεH with ε=1 , ε=0.5 , ε=0.25 ε=0.1 , ε=0.0001 (lines with colors ranging from light grey to dark grey) and the RFOU process (bold black line). Note that the purple path (ε=0.0001) is not visible on the plot since it almost completely coincides with the bold black trajectory of the RFOU process.

When ε=0.0001, the path of YεH is so close to the corresponding path of the ROU process (bold black) that they are not distinguishable on the plot.

Acknowledgments

The present research is carried out within the frame and support of the ToppForsk project nr. 274410 of the Research Council of Norway with title STORM: Stochastics for Time-Space Risk Models. The first author is supported by the National Research Fund of Ukraine under grant 2020.02/0026.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The present research is carried out within the frame and support of the ToppForsk project nr. ( Norges Forskningsråd) [grant number 274410] of the Research Council of Norway with title STORM: Stochastics for Time-Space Risk Models. The first author is supported by the National Research Fund of Ukraine [2020.02/0026].

References

  • O.O. Aalen and H.K. Gjessing, Survival models based on the Ornstein-Uhlenbeck process, Lifetime Data Anal. 10 (2004), pp. 407–423.
  • C.A. Ball and A. Roma, Stochastic volatility option pricing, J. Financ. Quantitative Anal. 29 (1994), pp. 589–607.
  • L. Bo, D. Tang, Y. Wang and X. Yang, On the conditional default probability in a regulated market: a structural approach, Quantitative Financ. 11 (2011), pp. 1695–1702.
  • L. Bo, Y. Wang and X. Yang, Some integral functionals of reflected SDEs and their applications in finance, Quantitative Financ. 11 (2011), pp. 343–348.
  • J.C. Cox, J.E. Ingersoll and S.A. Ross, A re-examination of traditional hypotheses about the term structure of interest rates, The J. Financ. 36 (1981), pp. 769–799.
  • J.C. Cox, J.E. Ingersoll and S.A. Ross, An intertemporal general equilibrium model of asset prices, Econometrica: J. Econ. Soc. 53 (1985), pp. 363.
  • J.C. Cox, J.E. Ingersoll and S.A. Ross, A theory of the term structure of interest rates, Econometrica: J. Econom. Soc. 53 (1985), pp. 385.
  • G. Di Nunno, Y. Mishura and A. Yurchenko-Tytarenko, Sandwiched SDEs with unbounded drift driven by Holder noises, preprint (2020), Available at ArXiv 2012.11465.
  • V. Giorno, A.G. Nobile and R. di Cesare, On the reflected Ornstein-Uhlenbeck process with catastrophes, Appl. Math. Comput. 218 (2012), pp. 11570–11582.
  • V. Giorno, A.G. Nobile and L.M. Ricciardi, On some diffusion approximations to queueing systems, Adv. Appl. Probab. 18 (1986), pp. 991–1014.
  • R.S. Goldstein and W.P. Keirstead, On the term structure of interest rates in the presence of reflecting and absorbing boundaries, preprint (1997). doi:10.2139/ssrn.19840.
  • S.L. Heston, A closed-form solution for options with stochastic volatility with applications to bond and currency options, Rev. Financ. Stud. 6 (1993), pp. 327–343.
  • J. Hong, C. Huang, M. Kamrani and X. Wang, Optimal strong convergence rate of a backward Euler type scheme for the Cox-Ingersoll-Ross model driven by fractional Brownian motion, Stoch. Process. Appl. 130 (2020), pp. 2675–2692.
  • N. Ikeda and S. Watanabe, A comparison theorem for solutions of stochastic differential equations and its applications, Osaka J. Math. 14 (1977), pp. 619–633.
  • N. Ikeda and S. Watanabe, Stochastic Differential Equations and Diffusion Processes, Elsevier, 1981.
  • I. Karatzas and S.E. Shreve, Brownian Motion and Stochastic Calculus, Springer, New York, 1998.
  • P.R. Krugman, Target zones and exchange rate dynamics, The Quarterly J. Econ. 106 (1991), pp. 669–682.
  • K. Kubilius, Estimation of the Hurst index of the solutions of fractional SDE with locally Lipschitz drift, Lithuanian Association of Nonlinear Analysts (LANA). Nonlinear Analysis. Modelling and Control 25 (2020), pp. 1059–1078.
  • C. Lee and J. Song, On drift parameter estimation for reflected fractional Ornstein-Uhlenbeck processes, Stoch. An Int. J. Probab. Stoch. Process. 88 (2016), pp. 751–778.
  • V. Linetsky, On the transition densities for reflected diffusions, Adv. Appl. Probab. 37 (2005), pp. 435–460.
  • Y. Maghsoodi, Solution of the extended CIR term structure and bond option valuation, Math. Financ. An Int. J. Math., Stat. Financial Econom. 6 (1996), pp. 89–109.
  • A. Melnikov, Y. Mishura and G. Shevchenko, Stochastic viability and comparison theorems for mixed stochastic differential equations, Methodol. Comput. Appl. Probab. 17 (2015), pp. 169–188.
  • Y. Mishura and A. Yurchenko-Tytarenko, Fractional Cox-Ingersoll-Ross process with non-zero “mean”, Modern Stoch.: Theor. Appl. 5 (2018), pp. 99–111.
  • V.I. Piterbarg, Twenty Lectures About Gaussian Processes, Atlantic Financial Press, 2015.
  • L.M. Ricciardi, Stochastic population theory: Diffusion processes, Mathematical Ecology, Springer, Berlin Heidelberg, (1986), pp. 191–238.
  • L.C.G. Rogers and D. Williams, Diffusions, Markov Processes and Martingales, Cambridge University Press, 2000.
  • R. Schöbel and J. Zhu, Stochastic volatility with an Ornstein-Uhlenbeck process: an extension, Rev. Financ. 3 (1999), pp. 23–46.
  • T. Shiga and S. Watanabe, Bessel diffusions as a one-parameter family of diffusion processes, Zeitschrift fur Wahrscheinlichkeitstheorie und Verwandte Gebiete 27 (1973), pp. 37–46.
  • A.V. Skorokhod, Stochastic equations for diffusion processes in a bounded region, Theor. Probab. Appl. 6 (1961), pp. 264–274.
  • A.R. Ward and P.W. Glynn, A diffusion approximation for a Markovian queue with reneging, Queueing Syst. 43 (2003), pp. 103–128.
  • A.R. Ward and P.W. Glynn, Properties of the reflected Ornstein-Uhlenbeck process, Queueing Syst.44 (2003), pp. 109–123.
  • A.R. Ward and P.W. Glynn, A diffusion approximation for a GI/GI/1 queue with balking or reneging, Queueing Syst. 50 (2005), pp. 371–400.
  • X. Yang, X. Li, G. Ren, Y. Wang and L. Bo, Modeling the exchange rates in a target zone by reflected Ornstein-Uhlenbeck process, SSRN Electron. J. (2012), 10.2139/ssrn.2107686.
  • D. Ylvisaker, A note on the absence of tangencies in Gaussian sample paths, Ann. Math. Stat. 39 (1968), pp. 261–262.
  • M. Zähle, Integration with respect to fractal functions and stochastic calculus. I, Probability Theory and Related Fields 111 (1998), pp. 333–374.
  • S.Q. Zhang and C. YuanStochastic differential equations driven by fractional Brownian motion with locally Lipschitz drift and their implicit Euler approximation, Proceedings of the Royal Society of Edinburgh: Section A Mathematics, (2020), pp. 1–27.
  • J. Zhu, Applications of Fourier Transform to Smile Modeling, Springer, Berlin Heidelberg, 2010.