• Nebyly nalezeny žádné výsledky

EULER-MARUYAMA APPROXIMATIONS IN MEAN-REVERTING STOCHASTIC VOLATILITY MODEL UNDER REGIME-SWITCHING

N/A
N/A
Protected

Academic year: 2022

Podíl "EULER-MARUYAMA APPROXIMATIONS IN MEAN-REVERTING STOCHASTIC VOLATILITY MODEL UNDER REGIME-SWITCHING"

Copied!
20
0
0

Načítání.... (zobrazit plný text nyní)

Fulltext

(1)

EULER-MARUYAMA APPROXIMATIONS IN MEAN-REVERTING STOCHASTIC VOLATILITY MODEL UNDER REGIME-SWITCHING

XUERONG MAO, AUBREY TRUMAN, AND CHENGGUI YUAN

Received 28 December 2005; Revised 9 February 2006; Accepted 9 February 2006

Stochastic differential equations (SDEs) under regime-switching have recently been de- veloped to model various financial quantities. In general, SDEs under regime-switching have no explicit solutions, so numerical methods for approximations have become one of the powerful techniques in the valuation of financial quantities. In this paper, we will concentrate on the Euler-Maruyama (EM) scheme for the typical hybrid mean-reverting θ-process. To overcome the mathematical difficulties arising from the regime-switching as well as the non-Lipschitz coefficients, several new techniques have been developed in this paper which should prove to be very useful in the numerical analysis of stochastic systems.

Copyright © 2006 Xuerong Mao et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

1. Introduction

In the well-known Black-Scholes model, the asset price is described by a geometric Brow- nian motion

dX(t)=μX(t)dt+νX(t)dw1(t), (1.1) wherew1(t) is a scalar Brownian motion,μis the rate of return of the underlying asset, andνis the volatility. In this classical model, Black and Scholes [2] assumed that the rate of return and the volatility are constants. However, it has been proved by many authors (see, e.g., [5,14,16,20]) that the volatility is itself an It ˆo process in many situations.

For instance, Hull and White [16] assume that the instantaneous varianceV=ν2obeys another geometric Brownian motion

dV(t)=αV(t)dt+βV(t)dw2(t), (1.2) whereα,βare constants whilew2(t) is another Brownian motion andw1(t) andw2(t)

Hindawi Publishing Corporation

Journal of Applied Mathematics and Stochastic Analysis Volume 2006, Article ID 80967, Pages1–20

DOI10.1155/JAMSA/2006/80967

(2)

have correlationρ. Heston [14] assumes that the varianceV obeys the mean-reverting square root process

dV(t)=αλV(t)dt+βV(t)dw2(t) (1.3) while the mean-reverting process

dV(t)=αλV(t)dt+βV(t)dw2(t) (1.4) is also proposed as the volatility process by others. In particular, Lewis [18] proposes the mean-revertingθ-process

dV(t)=αλV(t)dt+βVθ(t)dw2(t), (1.5) whereθ1/2. This process unifies processes (1.3) and (1.4).

On the other hand, the rate of returnμis not a constant either and there is a strong evidence to indicate that it is a Markov jump process (see, e.g., [4,6,7,10,17,22,23,25]).

Of course, when the rate jumps, the volatility will jump accordingly. For example, the hybrid geometric Brownian motion

dX(t)=μr(t)X(t)dt+νr(t)X(t)dw1(t) (1.6) has been proposed by several authors (see [27,28] among others). Here,r(t) is a Markov chain with a finite state spaceᏹ= {1, 2,...,N}andμ,νare mappings fromᏹto [0,).

Equation (1.6) is also known as the geometric Brownian motion under regime-switching.

We observe that in this model, the volatility is also assumed to obey a Markov jump pro- cess. Recalling the stochastic volatility models mentioned above, we may more reasonably assume that the volatility process obeys a stochastic differential equation (SDE) under regime-switching, for example, the hybrid mean-revertingθ-process

dV(t)=αr(t)λr(t)V(t)dt+βr(t)Vθ(t)dw2(t). (1.7) Such stochastic models under regime-switching have recently been developed to model various financial quantities, for example, option pricing [4,10–13,17], stock returns [6, 7,23], and portfolio optimization [22,25]. In particular, the mean-reverting square root process under regime-switching or, more generally, (1.7) has found its considerable use as a model for volatility and interest rate. In general, SDEs under regime-switching have no explicit solutions so the Monte Carlo simulations have become one of the powerful techniques in valuation of financial quantities, for example, option price (see [9,15,24]).

However, there is currently a lack of theory that guarantees the convergence of the Monte Carlo simulations for SDEs under regime-switching in finance. This is due to the fact that most of SDEs under regime-switching in finance are nonlinear and non-Lipschitzian so we cannot appeal to the standard convergence theory for numerical simulations, as typ- ified by [26], to deduce that the numerically computed paths are accurate for small step sizes.

In this paper, we will concentrate on the Euler-Maruyama (EM) scheme for the typical hybrid mean-revertingθ-process (1.7) but the theory established here can certainly be de- veloped to cope with other SDEs under regime-switching in finance. InSection 2, we will

(3)

introduce necessary notations and investigate the global positive or nonnegative solutions to the mean-revertingθ-process under regime-switching. The EM numerical scheme will be defined inSection 3, where we will explain how to simulate discrete Markov chains, and hence the EM approximate solutions. InSection 4, we will show that the EM so- lutions converge to the exact solution. The path-dependent option with the volatility described by the hybrid mean-revertingθ-process will be discussed inSection 5, while Section 6contains applications to other financial quantities.

2. Nonnegative solutions

Throughout this paper, we let (Ω,Ᏺ,{t}t0,P) be a complete probability space with a filtration{t}t0 satisfying the usual conditions (i.e., it is increasing and right contin- uous whileᏲ0contains allP-null sets). Letw(t) be a scalar Brownian motion defined on the probability space. Let| · |denote the Euclidean norm. Letr(t),t0, be a right- continuous Markov chain on the probability space taking values in a finite state space ᏹ= {1, 2,...,N}with the generatorΓ=i j)N×Ngiven by

P

r(t+δ)=j|r(t)=i=

γi jδ+o(δ) ifi=j,

1 +γi jδ+o(δ) ifi=j, (2.1) whereδ >0. Hereγi jis the transition rate fromitojandγi j>0 ifi=jwhile

γii= −

j=i

γi j. (2.2)

We assume that the Markov chainr(·) is independent of the Brownian motionw(·). It is well known that almost every sample path ofr(·) is a right-continuous step function with a finite number of sample jumps in any finite subinterval ofR+:=[0,).

Let 1/2θ1. Consider the mean-revertingθ-process under regime-switching of the form

dS(t)=λr(t)μr(t)S(t)dt+σr(t)Sθ(t)dw(t), t0, (2.3) with initial dataS(0)=S0>0 andr(0)=i0ᏹ. Hereλ(i),μ(i),σ(i),iᏹ, are positive constants. The initial dataS0andi0 could be random, but the Markov property ensures that it is sufficient to consider only the case when bothS0andi0are constants. We note that the case whenθ=1/2 and the state space of the Markov chainᏹ= {1}corresponds to the classical mean-reverting square root process (1.3) (without regime-switching).

Since (2.3) is mainly used to model stochastic volatility or interest rate or an asset price, it is critical that the solutionS(t) will never become negative. The following lemma reveals this nonnegative property.

(4)

Lemma 2.1. For given any initial dataS(0)=S0>0 andr(0)=i0ᏹ, the solutionS(t) of (2.3) will never become negative with probability 1.

Proof. Clearly, the statement of the lemma is equivalent to that the solution of equation

dS(t)=λr(t)μr(t)S(t)dt+σr(t)S(t)θdw(t), t0, (2.4) will never become negative with probability 1 for any initial dataS(0)=S0>0 andr(0)= i0ᏹ. To show this, leta0=1, and for each integerk=1, 2,...,

ak=

ek(k+1) ifθ=1 2, (2θ1)k(k+ 1)

2

1/(12θ)

if1

2 < θ1,

(2.5)

so that

ak1

ak

du

u2θ =k. (2.6)

For each k=1, 2,..., there clearly exists a continuous function ψk(u) with support in (ak,ak1) such that

0ψk(u) 2

ku2θ forak< u < ak1 (2.7) andaakk1ψk(u)du=1. Defineϕk(x)=0 forx0 and

ϕk(x)= x

0 dy y

0 ψk(u)du forx <0. (2.8) ThenϕkC2(R,R) and has the following properties:

(i)1ϕk(x)0 forak< x < ak1, or otherwiseϕk(x)=0;

(ii)|ϕk(x)| ≤2/k|x|2θforak< x < ak1, or otherwiseϕk(x)=0;

(iii)|x| −ak1ϕk(x)≤ |x|for allxR.

Let ¯λ=maxiλ(i), let ¯μ=maxiμ(i), and let ¯σ=maxiσ(i). Now for anyt0, by the well-known It ˆo formula (see [19,21]), we can derive that

EϕkS(t)=ϕkS0

+E

t

0

λr(u)μr(u)S(u)ϕkS(u)+σ2r(u)

2 S(u)2θϕkSr(u)du

σ¯2t k .

(2.9) Hence

ak1ES(t)ak1σ¯2t

k , (2.10)

(5)

whereS(t)= −S(t) ifS(t)<0, or otherwiseS(t)=0. Lettingk→ ∞, we get thatES(t)

=0 for allt0. This implies thatS(t)0 for allt0 with probability 1 as required.

Furthermore, the following lemma reveals the (strictly) positive property of the solu- tion.

Lemma 2.2. For given any initial dataS(0)=S0>0 andr(0)=i0ᏹ, the solutionS(t) of (2.3) will remain positive with probability 1, namelyS(t)>0 for allt0 almost surely, if one of the following two conditions holds:

(i) 1/2< θ1;

(ii)θ=1/2 andσ2(i)2λ(i)μ(i) for alli.

To show this lemma, let us first invoke the standard results, for example, results of G¯ıhman and Skorohod [8], to establish the following result.

Lemma 2.3. Consider the mean-revertingθ-process

dX(t)=λμX(t)dt+σXθ(t)dw(t) (2.11) ont0 with initial valueX(0)=x0>0, where 1/2θ1 andλ,μ,σ are all positive constants. Then

(i) with probability 1, the solutionX(t) takes an infinite time to reach origin 0 if either 1/2< θ1 orθ=1/2 withσ22λμ;

(ii) with positive probability, the solutionX(t) reaches the origin in finite time ifθ=1/2 andσ2>2λμ.

Proof. The coefficients of (2.11)

a(z)=λ(μz), b(z)=σzθ (2.12)

obey the linear growth condition onzR+, so the solution will never explode to infinity in any finite time with probability 1. We therefore need only to consider if it reaches the origin in finite time or not.

Consider

L1= 1

0exp

x

1

2a(z) b2(z)dz

dx. (2.13)

When 1/2< θ1, this gives L1=C1

1 0exp

2λμx1

σ2(12θ)+ 2λx22θ σ2(22θ)

dx, (2.14)

whereC1is a positive constant. A simple inspection shows thatL1diverges. Hence, the required result when 1/2< θ1 follows from G¯ıhman and Skorohod [8, Chapter 21].

Similarly, whenθ=1/2, L1=C2

1

0exp

2λμ σ2 log(x)

dx=C2

1

0x2λμ/σ2dx, (2.15)

(6)

whereC2is another positive constant. It is then easy to see thatL1= ∞if 2λμσ2while L1<if 2λμ < σ2. The required results corresponding to 2λμσ2 or 2λμ < σ2 when θ=1/2 follow from G¯ıhman and Skorohod [8, Chapter 21] again.

UsingLemma 2.3, we can now proveLemma 2.2.

Proof ofLemma 2.2. It is well known (see, e.g., [1]) that there is a sequence of stopping times 0=τ0< τ1<···< τk→ ∞such that the Markov chainr(t) has the representation

r(t)= k=0

rτk

Ikk+1)(t), t0, (2.16)

whereIAis the indicator function of setA. Hence, fort[0,τ1], (2.3) becomes dS(t)=λr0

μr0

S(t)dt+σr0

Sθ(t)dw(t) (2.17)

withS(0)>0. This is a mean-revertingθ-process of type (2.11). ApplyingLemma 2.3, we observe that S(t)>0 for all t[0,τ1] with probability 1 under the conditions of Lemma 2.2. Now, fort12], (2.3) becomes

dS(t)=λrτ1

μrτ1

S(t)dt+σrτ1

Sθ(t)dw(t) (2.18)

with initial valueS(τ1)>0 a.s. Again, this is a mean-revertingθ-process of type (2.11).

ByLemma 2.3, we see thatS(t)>0 for allt12] with probability 1. Repeating this procedure, we see thatS(t)>0 for allt0 with probability 1 as required.

It is still remaining open whether the solutionS(t) will reach the origin in finite time with positive probability in the case whenθ=1/2 whileσ2(i)2λ(i)μ(i) does not hold for alliᏹ.

3. The Euler-Maruyama method

To define the Euler-Maruyama approximate solution, we will need the following lemma (see [1]).

Lemma 3.1. GivenΔ >0, letrkΔ=r(kΔ) fork=0, 1, 2,....Then{rkΔ, k=0, 1, 2,...}is a discrete-time Markov chain with the one-step transition probability matrix

P(Δ)=

Pi j(Δ)N×N=eΔΓ. (3.1) Given a step size Δ >0, the discrete-time Markov chain {rkΔ, k=0, 1, 2,...} can be simulated as follows: compute the one-step transition-probability matrix

P(Δ)=

Pi j(Δ)N×N=eΔΓ. (3.2)

(7)

Letr0Δ=i0 and generate a random numberξ1 which is uniformly distributed in [0, 1].

Define

r1Δ=

i1 ifi1− {N}is such that

i11 j=1

Pi0,j(Δ)ξ1<

i1

j=1

Pi0,j(Δ), N if

N1 j=1

Pi0,j(Δ)ξ1,

(3.3)

where we set0i=1Pi0,j(Δ)=0 as usual. Generate independently a new random number ξ2which is again uniformly distributed in [0, 1] and then define

r2Δ=

i2 ifi2− {N}is such that

i21 j=1

Pr1Δ,j(Δ)ξ2<

i2

j=1

Pr1Δ,j(Δ), N if

N1 j=1

Pr1Δ,j(Δ)ξ2.

(3.4)

Repeating this procedure, a trajectory of{rkΔ,k=0, 1, 2,...}can be generated. This pro- cedure can be carried out independently to obtain more trajectories.

After explaining how to simulate the discrete-time Markov chain{rkΔ,k=0, 1,...}, we can now define the EM approximate solution for (2.3). Given a step sizeΔ >0, lettk= fork=0, 1, 2,.... Compute the discrete approximationsskS(tk) by settings0=S0,r0Δ= i0and forming

sk+1=sk+λrkΔμrkΔsk

Δ+σrkΔskθΔwk, k=0, 1, 2,..., (3.5) whereΔwk=w(tk+1)w(tk). Let

s(t)¯ =sk, r(t)¯ =rΔk fort tk,tk+1

,k=0, 1, 2,..., (3.6)

and define the continuous EM approximate solution by s(t)=s0+

t

0λr(u)¯ μr(u)¯ ¯s(u)du+ t

0σr(u)¯ s(u)¯ θdw(u). (3.7) Note thats(tk)=s(t¯ k)=sk, that is,s(t) and ¯s(t) coincide with the discrete approximate solution at the grid points.

4. Convergence of the EM approximate solution

Since the coefficients of (2.3) satisfy the linear growth condition, by [26], we have the following lemma.

Lemma 4.1. LetS(t) be the solution of (2.3). Then for anyp1, there is a constantK, which is dependent on onlyp,T,S0but independent ofΔ, such that the exact solution and the EM

(8)

approximate solution to (2.3) have the property that

E sup

0tT

S(t)p

E sup

0tT

s(t)p

K. (4.1)

From this follows easily the following useful result.

Lemma 4.2. There is a constantC, which is independent ofΔ, such that

Es(t)s(t)¯ 2θθ, t[0,T]. (4.2) Proof. From now on,Cused in the proofs below will be a generic positive number inde- pendent ofΔbut may have different values where it appears.

For anyt[0,T], letkt=[t/Δ], the integer part oft/Δ. ByLemma 4.1, we then derive that

Es(t)¯ s(t)24(¯λλ¯μ¯σ)¯ E

1 +skt2

Δ2+w(t)wktΔ2CΔ. (4.3) So, since 1/2θ1, by the Lyapunov inequality, we get

Es(t)¯ s(t)

Es(t)¯ s(t)2θθ (4.4)

as required.

We can now state one of our main results.

Theorem 4.3. For each integerk=1, 2,...,

sup

0tTES(t)s(t)eλT¯

ek(k1)/2+4¯σ2T k +

1 ka2kθ+ 1

θ+o(Δ)

, (4.5)

whereCis a constant which is independent of the step sizeΔand ¯λ, ¯σhave been defined in the proof ofLemma 2.1.

Proof. Note that

S(t)s(t)= t

0

λr(u)μr(u)λr(u)¯ μr(u)¯ λr(u)S(u) +λr(u)¯ s(u)¯ du

+ t

0

σr(u)S(u)θσr(u)¯ ¯s(u)θdw(u).

(4.6)

(9)

Letϕkbe the same as defined in the proof ofLemma 2.1. Applying the It ˆo formula gives EϕkS(t)s(t)

=E t

0ϕkS(u)s(u)λr(u)μr(u)λr(u)¯ μr(u)¯

λr(u)S(u) +λr(u)¯ s(u)¯ du +1

2E t

0ϕkS(u)s(u)σr(u)S(u)θσr(u)¯ s(u)¯ θ2du=:I(t) +1 2J(t).

(4.7) By property (i) ofϕk,

I(t)E t

0

ϕkS(u)s(u)λr(u)μr(u)λr(u)¯ μr(u)¯

λr(u)S(u) +λr(u)¯ s(u)¯ du

E t

0

λr(u)μr(u)λr(u)¯ μr(u)¯ du

+E t

0

λr(u)S(u)λr(u)¯ s(u)¯ du.

(4.8)

Letn=[T/Δ], the integer part ofT/Δ. Then E

T

0

λr(u)μr(u)λr(u)¯ μr(u)¯ du

=n

k=0

E tk+1

tk

λr(u)μr(u)λr(u)¯ μr(u)¯ du

(4.9)

withtn+1being now set to beT. LetIGbe the indicator function of setGand compute E

tk+1

tk

λr(u)μr(u)λr(u)¯ μr(u)¯ duλμE¯ tk+1

tk

I{r(u)=r(tk)}du

λμ¯ tk+1

tk

P

r(u)=rtk

du=λμ¯ tk+1

tk

i

P rtk

=iP

r(u)=i|rtk

=idu

=λμ¯ tk+1

tk

i

P

rtk=i

j=i

γi jutk+ouTk

λμ¯

1maxiN

γii

Δ+o(Δ)

Δ.

(4.10)

(10)

Therefore, E

T

0

λr(u)μr(u)λr(u)¯ μr(u)¯ duλμ¯

1maxiN

γiiΔ+o(Δ). (4.11)

On the other hand, E

t

0

λr(u)S(u)λr(u)¯ ¯s(u)du

E t

0

λr(u)λr(u)¯ s(u)¯ du+E t

0λr(u)S(u)s(u)¯ du.

(4.12)

But E

t

0

λr(u)λr(u)¯ s(u)¯ du

=n

k=0

tk+1

tk

E

Eλr(u)λrtksk|I{r(u)=r(tk)} du

= n k=0

tk+1

tk

E

Eλr(u)λrtk|I{r(u)=r(tk)}

Esk|I{r(u)=r(tk)}

,

(4.13)

where in the last step we use the fact thatskandI{r(u)=r(tk)}are conditionally independent with respect to theσ-algebra generated byr(tk). In the same way as in (4.10), we have

E t

0

λr(u)λr(u)¯ ¯s(u)du

λμ¯

1maxiN

γii

Δ+o(Δ) T

0 Es(u)¯ du.

(4.14)

So, byLemma 4.1, E

t

0

λr(u)λr(u)¯ s(u)¯ du2(1 +Kλμ¯

1maxiN

γiiΔ+o(Δ)

. (4.15) Substituting (4.15) into (4.12) and usingLemma 4.2, we obtain

E t

0

λr(u)S(u)λr(u)¯ ¯s(u)du

+o(Δ) + ¯λE t

0

S(u)s(u)¯ du

+o(Δ) + ¯λE t

0

S(u)s(u)du,

(4.16)

(11)

whereCis a positive constant independent of Δand it may change line by line. This, together with (4.11), yields

I(t)+o(Δ) + ¯λE t

0

S(u)s(u)du. (4.17)

In the following, we will estimateJ(t):

J(t)σ2E t

0

ϕkS(u)s(u)S(u)θ¯s(u)θ2du

+ 2E t

0

ϕkS(u)s(u)σr(u)σr(u)¯ 2S(u)du.

(4.18)

Using property (ii) ofϕkandLemma 4.2, we have E

t

0

ϕkS(u)s(u)S(u)θs(u)¯ θ2du

E t

0

ϕkS(u)s(u)S(u)¯s(u)2θdu

22θ1E t

0

ϕkS(u)s(u)S(u)s(u)2θdu + 22θ1E

t

0

ϕkS(u)s(u)s(u)s(u)¯ 2θdu

2E t

0

2

kI{ak<|S(u)s(u)|<ak1}du+ 2 t

0

2

kak Es(u)¯s(u)2θdu4T k +θ

kak . (4.19) In the same way as (4.15) was proved, we can show that

E t

0

ϕkS(u)s(u)σr(u)σr(u)¯ 2S(u)2θdu

E t

0

2 ka2kθ

σr(u)σr(u)¯ 2S(u)2θdu+o(Δ) ka2kθ .

(4.20)

Substituting (4.20) and (4.19) into (4.18), we have J(t)σ2T

k +θ+o(Δ)

ka2kθ . (4.21)

Therefore, Eϕk

S(t)s(t)σ2T

k +θ+o(Δ)

ka2kθ ++o(Δ) + 2¯λE t

0

S(u)s(u)du. (4.22)

(12)

Noting that

Eϕk

S(t)s(t)ES(t)s(t)ak1 (4.23)

gives

ES(t)s(t)ak1+4¯σ2T k +

1

ka2kθ + 1θ+o(Δ)+ ¯λ t

0ES(u)s(u)du.

(4.24) The required assertion follows finally from the Gronwall inequality.

Next, we derive a bound for a stronger form of the error. This version uses anL2- distance and places the supremum over time inside the expectation operation. The result below involves theL1-error which is explicitly bounded inTheorem 4.3, and hence is also computable.

Theorem 4.4. One has E

sup

0tT

S(t)s(t)2

e(8¯σ2+2¯λ2)T2

+o(Δ) + 8¯σ2T sup

0uTES(u)s(u)

. (4.25) Proof. For any 0tT, using the Cauchy-Schwarz inequality, we have

S(t)s(t)2

T t

0

λr(u)μr(u)λr(u)¯ μr(u)¯ λr(u)S(u) +λr(u)¯ s(u)¯ 2du

+ t

0

σr(u)S(u)θσr(u)¯ s(u)¯ θdw(u) 2

.

(4.26) In the same way as (4.11) and (4.16) were proved, we derive

E t

0

λr(u)μr(u)λr(u)¯ μr(u)¯ λr(u)S(u) +λr(u)¯ ¯s(u)2du

2E t

0

λr(u)μr(u)λr(u)¯ μr(u)¯ 2du

+E t

0

λr(u)S(u)λr(u)¯ ¯s(u)2du

+o(Δ) + 2¯λ2E t

0

S(u)s(u)2du.

(4.27)

(13)

Using the Doob martingale inequality (see [19]), we find that for anyt1[0,T], E

sup

0tt1

t

0

σr(u)S(u)θσr(u)¯ s(u)¯ θdw(u) 2

4E t1

0

σr(u)S(u)θσr(u)¯ s(u)¯ θ2du

+o(Δ) + 8¯σ2E t1

0

S(u)s(u)2θdu

+o(Δ) + 8¯σ2E t1

0

S(u)s(u)du+ 8¯σ2E t1

0

S(u)s(u)2du.

(4.28)

Therefore, E

sup

0tt1

S(t)s(t)2

+o(Δ) +σ2+ 2¯λ2E t1

0

S(u)s(u)2du+ 8¯σ2E t1

0

S(u)s(u)du

+o(Δ) +σ2+ 2¯λ2t

1

0 E

sup

0uv

S(u)s(u)2

dv + 8¯σ2T sup

0uTES(u)s(u).

(4.29) An application of the Gronwall inequality completes the proof.

5. Options under stochastic volatility and regime-switching

In this section, we study the Heston stochastic volatility model under regime-switching, namely

dX(t)=λ1

r(t)μ1

r(t)X(t)dt+σ1

r(t)X(t)V(t)dw1(t), (5.1) dV(t)=λ2

r(t)μ2

r(t)V(t)dt+σ2

r(t)Vθ(t)dw2(t), 0tT. (5.2) HereV(t) is the volatility that feeds into the asset priceX(t). The Brownian motions w1(t) andw2(t) may be correlated. Naturally, we assume that the initial valuesX(0) and V(0) are both positive constants. Moreover,λ1,σ1, and so forth are all mappings fromᏹ toR+.

We begin with a lemma showing that the positivity in the initial data leads to the positive solutionX(t).

Lemma 5.1. IfV(t),t[0,T], is given by (5.2), then P

X(t)>00tT=1. (5.3)

Odkazy

Související dokumenty

In this paper, the Monte Carlo method has been used for estimating the uncertainty of correlated in- put quantities by direct measurement of the sides of a rectangle using the

I study the effects of (i) the HH’s ability to monitor the government and (ii) the cost of changing the government on equilibrium outcomes: persistence of regimes and frequency

The results of the research show that the use of sports celebrities in marketing communication for generation Y does not have a proven effect on positive word of mouth or on

In this paper, we have performed the robustness and sensitivity analysis of several continuous-time stochastic volatility models (Heston, Bates and FSV model) with respect to

Yin, “Markowitz’s mean-variance portfolio selection with regime switching: a continuous-time model,” SIAM Journal on Control and Optimization, vol... Li,

In Section 3 we study the current time correlations for stationary lattice gases and in Section 4 we report on Monte-Carlo simulations of the TASEP in support of our

Analysis of a stochastic delay competition system driven by Lévy noise under regime switching.. Shiying Li and Shuwen

The joint switching seems to be stochastic and the switching may appear in any joints of the robot which need to develop Markovian jump linear system MJLS 37 to incorporate