Skip to main content

ORIGINAL RESEARCH article

Front. Appl. Math. Stat., 30 March 2022
Sec. Dynamical Systems
This article is part of the Research Topic Modelling and Numerical Simulations with Differential Equations in Mathematical Biology, Medicine and the Environment View all 11 articles

On the Global Positivity Solutions of Non-homogeneous Stochastic Differential Equations

  • Department of Mathematics and Applied Mathematics, University of Limpopo, Turfloop Campus, Sovenga, South Africa

In this article, we treat the existence and uniqueness of strong solutions to the Cauchy problem of stochastic equations of the form dXt=αXtdt+σXtγdBt,X0=x>0. The construction does not require the drift and the diffusion coefficients to be Lipschitz continuous. Sufficient and necessary conditions for the existence of a global positive solution of non-homogeneous stochastic differential equations with a non-Lipschitzian diffusion coefficient are sought using probabilistic arguments. The special case γ = 2 and the general case, that is, γ > 1 are considered. A complete description of every possible behavior of the process Xt at the boundary points of the state interval is provided. For applications, the Cox-Ingersoll-Ross model is considered.

1. Introduction

The theory of stochastic differential equations was developed by [1]. Stochastic differential equations are valuable tools for modeling systems and processes with stochastic disturbances in many fields of science and engineering. For the general theory of stochastic differential equations, one can refer to [25]. Several authors have discussed results concerning the existence and uniqueness of solutions of stochastic differential equations [2, 6, 7]. Mishura and Posashkova [8] provided a sufficient condition on coefficients which ensures almost surely positivity of the trajectories of the solution of the stochastic differential equation with non-homogeneous coefficients and non-Lipschitz diffusion. Appleby et al. [9] investigated highly non-linear stochastic differential equations with delays and showed that properties on the coefficients of stochastic differential equations that guarantee stability also guarantee positivity of solutions as long as the initial value is non-zero. Xu et al. [10] investigated the global positive solution of a stochastic differential equation, where they generalized the mean-reverting constant elasticity of variance process by replacing the constant parameters with the parameters modulated by a continuous-time, finite-state, Markov chain. Zhang [11] treated the properties of solutions to stochastic differential equations with Sobolev diffusion coefficients and singular drifts. Bae et al. [12] proved the existence of and uniqueness of solution to stochastic differential equations under weakened and Ho¨lder conditions and a weakened linear growth condition. Conditions for positivity of solutions of fractional stochastic differential equations with coefficients that do not satisfy the linear growth Lipschitz continuous conditions were obtained by [13].

The aim of this article is to prove the existence of a global positive solution of stochastic differential equations of the form

dXt=αXtdt+σXtγdBt,X0=x>0,    (1)

where Bt is a standard Brownian motion, for different values of γ where α denotes the drift, σ denotes the volatility. X = (Xt)t≥0 describes the underlying asset price. Such stochastic differential equations arise in modeling asset prices and interest rates on financial markets and it is crucial that Xt never becomes negative. Mao and Yuan [2] discussed the analytical properties when 12γ1 and showed that for a given initial value X0 = x > 0, the solution of (1) remains positive with probability 1, namely, Xt > 0 for all t ≥ 0 almost surely. The cases γ = 0 and γ = 1 give rise to the Ornstein-Uhlenbeck process and the Geometric Brownian motion, respectively, and this has been dealt with in the literature, see [4, 5, 14, 15]. When γ > 1, the diffusion coefficient of Equation (1) does not satisfy the linear growth condition, even though it is locally Lipschitz continuous. In view of this, it is not straightforward from the general theory of stochastic differential equations to obtain a unique global positive solution to Equation (1) that is defined for all t ≥ 0. Nevertheless, there is a way to overcome such difficulties which we present in this article and we also provide detailed proofs that there is unique solutions to equations of the form (1). This article is an extension of the work done in [16] and [17] to non-homogeneous stochastic differential equations.

This article is structured as follows. In Section 2, we consider the existence of a positive global solution for non-homogeneous stochastic differential equations with non-Lipschitz coefficients. In particular, we treat the case γ = 2 and prove that if α ≥ 0 and x ≥ 0 is arbitrary, then a unique strong solution of Equation (1) exists. Section 3 deals with the existence and uniqueness of a positive global solution to non-homogeneous stochastic differential equation (1). In particular, we consider the general case, that is, γ > 1. We provide a detailed proof of the existence of a unique solution to Equation (1). In Section 4, we investigate the behavior of the underlying process Xt at the boundaries of the state space (0, ∞). The main tool used are simple probabilistic arguments. We only require the coefficients of our model to be continuous in the usual sense. In Section 5, we provide a brief conclusion.

2. Existence of Positive Global Solutions: γ = 2

We want to prove that a unique global positive solution to Equation (1) exists and investigate its properties. We notice that if X0 = x = 0, then by strong uniqueness we have Xt = 0 for all t ≥ 0. In addition, if a solution Xt exists for all t < τ(ω) ≤ ∞ for some x > 0 and XT = 0 for some T = T(ω) < τ(ω), then by the Strong Markov property we haveXt = 0 for all t ∈ [T(ω), τ(ω)]. In particular, x ≥ 0 implies that Xt ≥ 0.

We call (Ω,F,(Ft),) a stochastic basis if (Ω,F,) is a complete probability space and (Ft) is a right continuous filtration on Ω augmented by the ℙ-null sets. Let B = (Bt)t≥0 be a standard Brownian motion defined on a stochastic basis (Ω,F,(Ft),).

We consider a stochastic differential equation of the form

dXt=α(Xt)dt+σ(Xt)dBt,    (2)

where the coefficients α:RR and σ:RR are both Borel measurable functions. By the definition of stochastic differential, Equation (2) is equivalent to the stochastic integral equation:

Xt=x+0tα(Xs)ds+0tσ(Xs)dBs.    (3)

Definition 2.1. [2, p. 48] An R-valued stochastic process {Xt}t ∈ [0, T] is called a solution of Equation (2) if it has the following properties:

1. {Xt} is continuous and Ft-adapted.

2. 0tα(Xs)ds< and 0tσ2(Xs)ds<.

3. Equation (3) holds for every t ∈ [0, T] with probability 1.

Definition 2.2. [18, p. 167] A solution (X, B) of Equation (2) defined on (Ω,F,(Ft),) is said to be a strong solution if X is adapted to the filtration {FtB}, that is, the filtration of B = (Bt)t≥0 completed with respect to ℙ.

Definition 2.3. [7, p. 300] A weak solution is a triple ((Ω,F,),(B,X),(Ft)) where (Ω,F,) is a probability space, {Ft} is a filtration of sub-σ-fields of F satisfying the usual conditions, X is a continuous, adapted R-valued process, B is the standard Brownian motion such that Equation (3) is satisfied.

Remark 2.1.

1. Definition 2.2 says that if the probability space (Ω,F,), the filtration {Ft}t0, the Brownian motion Bt and the coefficients α(x) and σ(x) are all given in advance, and then the solution Xt is constructed, such a solution is called a strong solution.

2. Definition 2.3 says that if we are only given the coefficients α(x) and σ(x), and we are allowed to construct a suitable probability space, a filtration and find a solution to the Equation (2), then such a solution is called a weak solution.

A solution {Xt}t∈[0,T] is said to be unique if any other solution {X¯t} is indistinguishable from {Xt}, that is

(Xt=X¯t     0tT)=1.

Following [19], we impose the following hypothesis:

(H) The drift coefficient is globally Lipschitz, that is, for all x, yR,

α(x)-α(y)Kx-y    (4)

where K is a fixed constant, while the diffusion coefficient is globally Ho¨lder continuous, that is, for all x, yR,

σ(x)-σ(y)h(x-y)    (5)

where h:[0, ∞) → [0, ∞) is a strictly increasing function with h(0) = 0 and the integral

0εduh2(u)=,   ε>0.

It is known [see [19]] that under the hypothesis (H), the strong uniqueness solution holds for the stochastic differential equation (2).

For the case γ = 2, Equation (1) takes the form

dXt=αXtdt+σXt2dBt.    (6)

If we let Yt = ln∣Xt∣ then an application of Itô's formula yields

dYt=(α-12σ2Xt2)dt+σXtdBt    (7)

which is equivalent to

ln Xt=ln x+αt-12σ20tXs2ds+σ0tXsdBs.    (8)

This solution presents a challenge as the coefficients in Equation (8) do not satisfy the linear growth and Lipschitz conditions. However, there is a way to go around this. In the next result we prove the existence of a global solution to Equation (6) following arguments presented in [20].

Theorem 2.1. Suppose α ≥ 0 and x ≥ 0 is arbitrary, then the stochastic differential equation of the form (6) has a unique, strong solution Xt defined for all t ≥ 0.

Proof: The result is proved by a truncation procedure. For each n ≥ 1, we set α = αn(x) and the truncation function

σn(x)={σx2if xn,σn2if x>n.

Then, αn(x) and σn(x) satisfy the hypothesis (H). Hence, there is a unique solution Xt=Xt(n) defined for all t to the equation

Xt(n)=x+0tαn(Xs(n))ds+0tσn(Xs(n))dBs.    (9)

Define the stopping time

τn=inf{t>0;Xt(n)n},  n=1,2,....    (10)

Then, by strong uniqueness we have

Xt(n)(ω)=Xt(n+1)(ω)  for all  tτn  a.s.    (11)

Therefore,

τn=inf{t>0;Xt(n)n}<inf{t>0;Xt(n+1)n+1}=            τn+1.    (12)

Hence, {τn} is an increasing sequence of stopping times. Put

τ(ω)=limnτn(ω).

Then, for t < τ(ω), a process Xt can be defined by setting

Xt(ω)=Xt(n)(ω),    if    t<τn(ω).    (13)

It is clear that if t < τ(ω) then one can easily show that t < τn(ω) for some n. Therefore, by (11), this defines Xt(ω) uniquely. Hence, we have

Xt=x+0tαXsds+0tσXs2dBs    for  t<τ(ω).    (14)

3. Existence and Uniqueness of Positive Global Solution

In this section, we provide a detailed proof that there is a unique positive global solution to Equation (1). In particular, we focus on the case γ > 1. To establish the existence of a unique positive global solution, we need the following result.

Lemma 3.1. [3, p. 57] The coefficients of Equation (1) satisfy the local Lipschitz condition for given initial condition X0 = x > 0, that is, for every integer k > 1, there exists a positive constant Lk and x, y ∈ [0, k] such that

αx-αy2+σxγ-σyγ2Lkx-y2.    (15)

Therefore, there exists a unique local solution to Equation (1).

We now state our result in the following theorem.

Theorem 3.1. For any given initial value X0 = x > 0, α and σ > 0 there exists a unique positive global solution Xt to Equation (1) on t ≥ 0 for γ > 1.

Proof: It is clear that the coefficients of Equation (1) are locally Lipschitz continuous. Therefore, for any given initial value X0 = x > 0, there is a unique local solution Xt, t ∈ [0, τ(ω)] of Equation (1) where τ(ω) is the explosion time. To prove that the solution is global, it suffices to show that τ(ω) = ∞ almost surely. We prove this by contradiction. If τ(ω) ≠ ∞, then we can find a pair of positive constants ϵ and T such that

(τ(ω)T)>ϵ.    (16)

For each integer n > 1, we define a stopping time

τn=inf{t0:Xtn}.    (17)

Since τn → τ(ω) almost surely, we can find a sufficiently large n0 for which

(τnT)>ϵ2,   for all nn0.    (18)

For θ, β > 0, we define a function VC2 as

V(X):=θX+βX-2,    (19)

which is continuously twice differentiable in X. We observe that V(X) → +∞ as X → +∞ or X → 0. For any 0 < t < T, an application of Ito^ formula gives

dV(Xt)=LV(Xt)dt+σXγ(12θXt-12-2βXt-3) dBt,    (20)

where

LV(Xt)=αXt(12θXt-12-2βXt-3)                     +12σ2Xt2γ(-14θXt-23+6βXt-4).    (21)

By boundedness of polynomials, there exists a constant K such that

αXt(12θXt-12-2βXt-3)+12σ2Xt2γ(-14θXt-23+6βXt-4)K.    (22)

Therefore, for any t ∈ [0, T]

E[V(Xtτn)]=V(x)+E[0tτnLV(Xs)ds]V(x)+KT                                +KE[0tE[V(Xsτn)]ds].    (23)

The application of the Grownwall inequality yields

E[V(XTτn)][V(x)+KT]eKT    (24)

which is equivalent to

E[V(Xτn)1{τnT}][V(x)+KT]eKT.    (25)

On the other hand, we define

Mn=inf{V(Xt)Xt>n,t[0,T]}.    (26)

As n → +∞, Mn → +∞. It now follows from (18) and (26) that

[V(x)+KT]eKTMn({τnT})12ϵMn.    (27)

Letting n → +∞ yields a contradiction, so we must have τ(ω) = ∞ almost surely. Therefore, there exists a unique positive global solution Xt to Equation (1) for all t ≥ 0.

4. Analysis of the Solution at the Boundaries of the State Space

We now investigate the behavior of the underlying process Xt at the boundaries of the state space (0, ∞) using probability arguments. Xt is the solution of the stochastic differential equation (1), where Xt is defined on the state space (0, ∞), that is, the whole positive real line.

We first consider the Ito^ diffusion of the form

dXt=α(Xt)dt+σ(Xt)dBt, X0=x,    (28)

where α:RR and σ:RR are functions satisfying the hypothesis (H). Note that here we do not have the time argument. We assume that the state space of Xt is a finite or infinite interval. Such a process is a continuous Markov process and under weak regularity conditions the drift coefficient α(x) and the diffusion coefficient σ(x) are characterized, respectively, by

α(x)=limh0h-1E[Xh-x]    (29)

and

σ2(x)=limh0h-1E[(Xh-x)2]=limh0h-1Var(Xh).    (30)

For details about these conditions as well as the foregoing, see [21]. The above conditions can conveniently be weakened to give the following three conditions.

h-1E[(Xh-x)1{|Xh-x|1}]α(x),    (31)
h-1E[(Xh-x)21{|Xh-x|1}]σ2(x),    (32)

and

h-1(|Xh-x|>ε)0    ε>0,    (33)

where 1{·} is the indicator function. These conditions enable us to perform the analysis of (29) without assuming the Lipschitz conditions to the coefficients. We will, however, assume that α(x) and σ(x) are continuous.

Fix qR and define the scale function u by

u(x)=qxexp(-qt2α(z)σ2(z)dz)dt,  u(q)=0.    (34)

The function u has a continuous, strictly positive derivative and u″ exists almost everywhere and satisfies

u(x)=-2α(x)σ2(x)u(x).    (35)

We also introduce the speed measure

m(dx)=2u(x)σ2(x)dx.    (36)

Now, let p(t, x, y) be the transition density of Xt. Then, the Kolmogorov backward equation is given by

pt=12σ2(x)2px2+α(x)px.    (37)

At t = 0, p(0, x, y) = δ(xy), where δ(·) is Dirac's delta function.

Let [a, b] be a fixed interval and start the process at X0 = x ∈ (a, b). We want to find the probability p+(x) that the process Xt hits b before it hits a. By the Markov property, we have

p+(x)=E[p+(Xs)]+O((|Xs-x|>ε)).

It follows from Equation (33) that

s-1((|Xs-x|>ε))0,

when s ↓ 0 if a + ε < x < b − ε. Using the Itô's formula and Equation (37), we can show that p+(x) satisfies the Kolmogorov's backward equation

12σ2(x)p+(x)+α(x)p+(x)=0,    (38)

for x ∈ (a, b) with the boundary conditions p+(a) = 0 and p+(b) = 1. The explicit solution to Equation (38) is

p+(x)=Aqxexp(-qt2α(z)σ2(z)dz)dt+B.    (39)

We can write Equation (39) in the form

p+(x)=Au(x)+B,    (40)

where u(x) is of the form Equation (34) for a fixed q ∈ (a, b), with A and B constants. Now, an application of boundary conditions p+(a) = 0 and p+(b) = 1 gives:

A=1u(b)-u(a)   and B=-u(a)u(b)-u(a),

so that

p+(x)=u(x)-u(a)u(b)-u(a).    (41)

Equations (34) and (41) will be important when applied to our specific problem.

Following similar arguments, we define

e(x)=E[Tab],    (42)

where Tab = inf{t > 0:Xt ∉ (a, b)}. An application of the Markov property gives

e(x)=s+E[e(Xs)]+O((|Xs-x|>ε)).

Dividing by s and letting s tend to 0 and an application of the Itô formula gives

12σ2(x)e(x)+α(x)e(x)=-1.    (43)

This equation can be solved by the standard Green function techniques as follows. The corresponding homogeneous equation is

e(x)+2α(x)σ2(x)e(x)=0

and its solution is

e(x)=u(x)-u(a)u(b)-u(a),    (44)

with boundary conditions e(a) = 0 and e(b) = 1 where u(x) is defined in Equation (34). The Green function, G(a,b)(x, y), is calculated as

G(a,b)(x,y)={1W·e1(x)e2(y)ifxy,1W·e1(y)e2(x)ifxy,    (45)

where e1 and e2 take the form of Equation (44) and W is the Wronskian given by

W=u(x)u(b)-u(a).

Therefore, the solution to Equation (43) is given by

e(x)=abG(a,b)(x,y)m(dy),

where G(a, b)(x, y) is given by

G(a,b)(x,y)={2(u(x)-u(a))(u(b)-u(y))u(b)-u(a) if  xy,2(u(y)-u(a))(u(b)-u(x))u(b)-u(a) if  x>y,    (46)

and m(dy) is given by Equation (36).

We now consider Equation (1). We note that the diffusion coefficient σ(x) = σxγ in Equation (1) is defined only on (0, ∞), that is, the state space of the process is made up of the positive reals. The process Xt in Equation (1) is a diffusion process, and the coefficients σ and α are continuous on (0, ∞). Following the arguments in [17], we investigate the behavior of Equation (1) at the boundaries of the state space. It is of interest whether or not the boundary points 0 and/or ∞ can be reached by the process Xt in a finite time.

A boundary point is said to be accessible if it can be reached in finite time with positive probability. Otherwise it is inaccessible [17]. The accessible boundary points are of two different types, namely, the exit and regular boundary points. For the exit boundary, the process is absorbed after the boundary is reached while the regular boundary point is imposed on a standard Brownian motion and could either be absorbed or reflected once the boundary is reached. The inaccessible boundaries are also of two types, namely, the entrance and natural boundary points. The boundary is said to be of entrance type if it is possible to start the process at infinity and then reach the interior of the state interval, otherwise it is called natural.

Let [a, b] be a fixed interval and the process Xt starts in X0 = x ∈ (a, b). Let α and σ be continuous on a state interval whose interior is (c, d). We note that we may have c = −∞ and/or d = +∞. It is assumed that σ2(x) > 0 on (c, d). Further, let (c~,d~)=(u(c),u(d)), where u is the scale function given by Equation (34).

Definition 4.1. A natural upper boundary point d is said to be attracting if there is a positive probability that Xt shall converge to d as t → ∞.

The following classification theorem, taken from [17], will be the framework of the analysis of Equation (1).

Theorem 4.1. Let u be the scale function given by Equation (34) and m(dy) be the speed measure given by Equation (36). Let b be a point in the interior of the state space (c, d). Then, the following statements hold.

1. A necessary and sufficient condition for d to be accessible is that u(d) < ∞ and bd(u(d)-u(y))m(dy)<.

2. An accessible boundary point d is regular if and only if bdm(dy)<. Otherwise it is exit boundary.

3. An inaccessible boundary point d is natural if and only if bdu(y)m(dy)=.

4. A natural boundary point d is attracting if and only if u(d) < ∞ and at the same time bdm(dy)=.

We are now in a position of analyzing the non-homogeneous stochastic differential equation (1), repeated here for convenience,

dXt=αXtdt+σXtγdBt,  γ>1, X0=x>0.    (47)

This is a diffusion process with α(x) = αx, σ(x) = σxγ and natural state interval c = 0 to d = ∞. Let b be a point in the interior of this state interval. From Equation (34) and (36), we calculate the scale function and speed measure, respectively, corresponding to Equation (47) to be

u(x)=σ22α(b2γ-1-x2γ-1exp(ασ2(1-γ)(b2-2γ-x2-2γ))),

and

m(dy)=1σ2y2γexp(-ασ2(1-γ)(b2-2γ-y2-2γ))dy.    (48)

It remains only to classify our boundary points on the basis of these results. We note that

u(d)=σ22α(b2γ-1-d2γ-1exp(ασ2(1-γ)(b2-2γ-d2-2γ))).

Since d = ∞ in our state space (0, ∞), we use a limiting argument:

limdu(d)=σ2b2γ-12α,

provided 0<γ<12. Now, since b is a finite fixed point in (0, ∞), the limit is finite. We also need to investigate the integral

bd(u(d)-u(y))m(dy)=bd12αydy-12αexp(-ασ2(1-γ)d2-2γ)d2γ-1     ×bd1y2γexp(ασ2(1-γ)y2-2γ)dy.    (49)

We observe that as d → ∞ the term exp(-ασ2(1-γ)d2-2γ) approaches 0 provided 0 < γ < 1. So in this case we remain with

bd(u(d)-u(y))m(dy)bd12αydy  as  d.

We therefore, according to Theorem 4.1, conclude that the upper boundary point d = ∞ is inaccessible if 0 < γ < 1. Now for the lower boundary point 0 we have

u(0)=σ2b2γ-12α<,

since b is a finite fixed point in the state space (0, ∞). We also have

0b(u(0)-u(y))m(dy)=0b12αydy,  as y0.

This shows that the boundary point 0 is inaccessible for all γ ≠ 1. It remains to establish whether our boundary points are natural or not. Theorem 4.1 says the boundary point d is natural if and only if

bdu(y)m(dy)=.

Now,

    bdu(y)m(dy)=    bd(b2γ-12αy2γexp(-ασ2(1-γ)(b2-2γ-y2-2γ))-12αy)dy=b2γ-1exp(-ασ2(1-γ)b2-2γ)2αbd1y2γexp(ασ2(1-γ)y2-2γ)dy    -bd12αydy.

We observe that for 0 < γ < 1 the integral

bd1y2γexp(ασ2(1-γ)y2-2γ)dy

explodes to infinity very fast as d → ∞. Although the second integral,

bd12αydy,

also tends to infinity as d → ∞, the whole integral bdu(y)m(dy) tends to infinity as d → ∞ because the second integral goes to infinity very slowly as compared to the first one. Hence,

bdu(y)m(dy)=,

provided 0 < γ < 1. This tells us that the boundary point d = ∞ is natural if 0 < γ < 1.Using similar arguments, we can show that

0bu(y)m(dy)=.

Therefore, for 0 < γ < 1, the boundary point 0 is natural. Next, we investigate if our natural boundary points are attracting. According to Theorem 4.1 the boundary point d is attracting if and only if u(d) < ∞ and at the same time bdm(dy)=. Now we have already seen that u(d) < ∞ if d = 0 and/or d = ∞ provided 0 < γ < 1. Further

bdm(dy)=1σ2exp(-ασ2(1-γ)b2-2γ)                          ×bd1y2γexp(ασ2(1-γ)y2-2γ)dy,

which, for the reason given before, explodes to infinity as d → ∞ for all 0 < γ < 1. Therefore,

bm(dy)=,

for 0 < γ < 1. Hence, the upper boundary point d = ∞ is attracting for 0 < γ < 1. Similarly the lower boundary point 0 is shown to be attracting.

Now, we have established that both boundary points are attracting when 0 < γ < 1. In this case we will show that, by Equation (41), our process will converge to ∞ with probability p+(x), where x = X0 ∈ (0, ∞). It turns out that

p+(x)=0xexp(-ασ2(1-γ)y2-2γ)dy0exp(-ασ2(1-γ)y2-2γ)dy,  0<γ<1.    (50)

Evaluating the integrals yields

p+(x)=limy(xy)2γ-1exp(ασ2(1-γ)(y2-2γ-x2-2γ))=0,

for 12<γ<1. This shows that although the upper boundary d = ∞ is attracting for 0 < γ < 1, the process Xt will not converge to ∞ if 12<γ<1. Furthermore, the process Xt converges to 0 with probability 1−p+(x) which turns out to be 1 in this case. Thus it is certain that Xt will converge to 0 when 12<γ<1. We observe that if 0<γ<12 we have a problem since, in this case, it is not possible to proceed using a probabilistic argument. Our analysis is not complete if we do not consider the case γ > 1. We now proceed to make this analysis. As seen earlier

u(d)=σ22α(b2γ-1-d2γ-1exp(ασ2(1-γ)(b2-2γ-d2-2γ))).

If γ > 1, for example, if γ = 2, we have

u(d)-  as  d,

since α, σ and b are fixed positive numbers. Therefore, we have

limdu(d)<       γ>1.

Observe also that for such γ we have that

u(0)=σ2b2γ-12α<,

effectively. Now, we consider again Equation (49). If γ > 1, for instance, γ = 2, we have

bd(u(d)-u(y))m(dy)=bd12αydy-d32αeασ2d2bd1y4e-ασ2y2dy.

We immediately observe that as d → ∞, bd(u(d)-u(y))m(dy)- since the integral bd12αydy goes to ∞ very slowly. Therefore, effectively we have

bd(u(d)-u(y))m(dy)<,

for the boundary point d = ∞ and whenever γ > 1. This, together with u(d) < ∞ for d = ∞ and γ > 1, shows that the upper boundary point d = ∞ is always accessible whenever γ > 1. However, it is clear that if d = 0,

0b(u(0)-u(y))m(dy)=0b12αydy=,

which shows that the lower boundary point 0 is always not accessible for γ > 1. In fact, the boundary point 0 is always inaccessible for all γ ≠ 1 as also shown earlier. So from definition we have seen that the upper boundary point ∞ can be reached in finite time with positive probability provided γ > 1.

Finally we want to classify the accessible boundary point ∞, that is. is it regular or exit? From Theorem 4.1 it is regular if and only if bdm(dy)<. Now, as obtained earlier on

bdm(dy)=1σ2exp(-ασ2(1-γ)b2-2γ)                          ×bd1y2γexp(ασ2(1-γ)y2-2γ)dy.

If γ > 1, the integral is always less than ∞ because of the negative exponent since α, σ are fixed positive numbers. So we have

bdm(dy)<   γ>1

since in this case the exponent is always negative. Hence for γ > 1, the accessible upper boundary point ∞ is always regular, i.e., apart from absorption and reflection there are also other possibilities after the boundary point is reached.

We, therefore, have the following result.

Theorem 4.2. Let x ∈ (0, ∞) with α ∈ R arbitrary. Then, the stochastic differential equation (1) has a unique strong global solution Xt:t ≥ 0. The solution has the following properties:

1. x = 0 implies that Xt = 0 for all t > 0 and x ≥ 0 implies Xt > 0 for all t ≥ 0.

2. If 12<γ<1, then limtXt=0 with probability 1 − p+(x) where p+(x) is given by Equation (50).

3. If γ > 1, then limtXt= with positive probability.

4. If γ = 1, we have the usual Geometric Brownian motion whereas if γ = 0, we have the Ornstein-Uhlenbeck process.

In mathematical finance, our result is of particular interest for the Cox-Ingersoll-Roll (CIR) model which describes the stochastic evolution of interest rates (rt)t≥0 by the stochastic differential equation

drt=α(μ-r)dt+σrtdBt,  t0,

with r0 ≥ 0 and αμ12σ2 where α, μ and σ denote real constants.

5. Concluding Remarks

In this article, we proved the existence of global positive solutions to non-homogeneous stochastic differential equations whose diffusion coefficient is non-Lispchitz. We relied on both the classical sense and probabilistic arguments. We provided detailed proofs in both cases. The probability arguments save as an alternative method of dealing with non-homogeneous stochastic differential equations where classical methods cannot be applied. Using the scale function and the speed of measure, we provided a complete classification of boundary types and boundary behavior of Equation (1). The results of this article can be applied to Cox-Ingersoll-Ross model. In addition, the positivity of solutions is important to other non-linear models that arise in sciences and engineering.

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary materials, further inquiries can be directed to the corresponding author.

Author Contributions

FM contributed to the conceptualization of the idea for research, development of the methodology, analysis of the model and writing up of the final article. LR conducted the primary research, identified the problem, and discussed the results. Both authors proof read the manuscript and approved the submitted version.

Funding

This work was supported by a Newton Fund's Operational Development Assistance grant, ID 32, under the SA-UK partnership. The grant is funded by the UK Department for Business, Energy and Industrial Strategy and Department of Higher Education and Training (DHET) and delivered by the British Council. For further information, please visit www.newtonfund.ac.uk.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

The authors are grateful to the referees for constructive comments and suggestions that have improved this article.

References

1. Itô K. On Stochastic Differential Equations. New York, NY: American Mathematical Society (1951).

Google Scholar

2. Mao X, Yuan C. Stochastic Differential Equations with Markovian Switching. London: Imperial College Press (2006). doi: 10.1142/p473

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Mao X. Stochastic Differential Equations and Applications. London: Elsevier (2007). doi: 10.1533/9780857099402

CrossRef Full Text | Google Scholar

4. Øksendal B. Stochastic Differential Equations: An Introduction With Applications. New York, NY: Springer Science & Business Media (2013).

PubMed Abstract | Google Scholar

5. Ikeda N, Watanabe S. Stochastic Differential Equations and Diffusion Processes. North-Holland: Elsevier (2014).

Google Scholar

6. Protter PE. Stochastic differential equations. In: Rozavskii B, Yor M, editors. Stochastic Integration and Differential Equations. New York, NY: Springer (2005). p. 249–361. doi: 10.1007/978-3-662-10061-5_6

CrossRef Full Text | Google Scholar

7. Karatzas I, Shreve S. Brownian Motion and Stochastic Calculus. vol. 113. New York, NY: Springer Science & Business Media (2012).

Google Scholar

8. Mishura Y, Posashkova S. Positivity of solution of nonhomogeneous stochastic differential equation with non-lipschitz diffusion. Theory Stochast Process. (2008) 14:77–88. Available online at: http://tsp.imath.kiev.ua/files/172/1434_7.pdf

Google Scholar

9. Appleby JA, Kelly C, Mao X, Rodkina A. Positivity and stabilisation for nonlinear stochastic delay differential equations. Stochastics Int J Probabil Stochast Process. (2009) 81:29–54. doi: 10.1080/17442500802214097

CrossRef Full Text | Google Scholar

10. Xu R, Wu D, Yi R. Existence theorem for mean-reverting CEV process with regime switching. In: 2015 International Conference on Mechatronics, Electronic, Industrial and Control Engineering (MEIC-15). Atlantis Press (2015). p. 1560–3. doi: 10.2991/meic-15.2015.357

CrossRef Full Text | Google Scholar

11. Zhang X. Stochastic differential equations with Sobolev diffusion and singular drift and applications. Ann Appl Probabil. (2016) 26:2697–732. doi: 10.1214/15-AAP1159

CrossRef Full Text | Google Scholar

12. Bae MJ, Park CH, Kim YH. An existence and uniqueness theorem of stochastic differential equations and the properties of their solution. J Appl Math Inform. (2019) 37:491–506. doi: 10.14317/jami.2019.491

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Kubilius K, Medziunas A. Positive solutions of the fractional SDEs with non-Lipschitz diffusion coefficient. Mathematics. MDPI (2021) 9:1–14. doi: 10.3390/math9010018

CrossRef Full Text | Google Scholar

14. Duffie D. Dynamic Asset Pricing Theory. Princeton, NJ: Princeton University Press (2010).

Google Scholar

15. Glasserman P. Monte carlo methods in financial engineering. In: Stochastic Modelling and Applied Probability. New York, NY: Springer-Verlag (2003). doi: 10.1007/978-0-387-21617-1

CrossRef Full Text | Google Scholar

16. Rundora L. Extension(s) of the Geometric Brownian Motion Model for Pricing of Assets. Masters Dissertation, Harare: University of Zimbabwe (1997).

17. Helland I. One-Dimensional Diffusion Processes and Their Boundaries. (1996).

Google Scholar

18. Revuz D, Yor M. Continuous Martingales and Brownian Motion. vol. 293. New York, NY: Springer Science & Business Media (2013).

Google Scholar

19. Yamada T, Watanabe S. On the uniqueness of solutions of stochastic differential equations. J Math Kyoto Univ. (1971) 11:155–67. doi: 10.1215/kjm/1250523691

CrossRef Full Text | Google Scholar

20. Lungu E, Øksendal B. Optimal harvesting from a population in a stochastic crowded environment. Math Biosci. (1997) 145:47–75. doi: 10.1016/S0025-5564(97)00029-1

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Gualtierotti A. Statistics for stochastic processes: applications to engineering and finance. In: Dodge Y, editor. Statistical Data Analysis and Inference. North-Holland: Elsevier (1989). p. 543–55. doi: 10.1016/B978-0-444-88029-1.50054-X

CrossRef Full Text | Google Scholar

Keywords: geometric Brownian motion, Itô diffusion, Lipschitz continuous, scale function, speed measure

Citation: Mhlanga FJ and Rundora L (2022) On the Global Positivity Solutions of Non-homogeneous Stochastic Differential Equations. Front. Appl. Math. Stat. 8:847896. doi: 10.3389/fams.2022.847896

Received: 03 January 2022; Accepted: 21 February 2022;
Published: 30 March 2022.

Edited by:

Ramoshweu Solomon Lebelo, Vaal University of Technology, South Africa

Reviewed by:

Yubin Yan, University of Chester, United Kingdom
Eduardo S. Zeron, Instituto Politécnico Nacional de México (CINVESTAV), Mexico

Copyright © 2022 Mhlanga and Rundora. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Farai Julius Mhlanga, farai.mhlanga@ul.ac.za

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.