θ | κ | σ | ρ | r | T |
0.09 | 2 | 0.375 | −0.5 | 0.015 | 0.087 |
In this paper a finite difference method (FDM) is provided for pricing perpetual timer options under the Heston volatility model. Considering the degeneracy of the pricing equation, we first prove the existence and uniqueness of the solution of the pricing problem with a new notion of boundary conditions at degenerate boundary and the infinity. Then we discuss the choice of artificial boundary value conditions and obtain a prior estimate of the internal error caused by the boundary value error. This estimate helps to choose appropriate artificial boundary values and solution domain to reduce internal error of the numerical solution. Furthermore, We build a FDM with second-order convergence for the pricing problem. Finally, we implement our method and show the visualization results.
Citation: Yaoyuan Zhang, Lihe Wang. Pricing perpetual timer options under Heston Model by finite difference method: Theory and implementation[J]. AIMS Mathematics, 2023, 8(7): 14978-14996. doi: 10.3934/math.2023764
[1] | Sun-Yong Choi, Donghyun Kim, Ji-Hun Yoon . An analytic pricing formula for timer options under constant elasticity of variance with stochastic volatility. AIMS Mathematics, 2024, 9(1): 2454-2472. doi: 10.3934/math.2024121 |
[2] | Min-Ku Lee, Jeong-Hoon Kim . Closed-form approximate solutions for stop-loss and Russian options with multiscale stochastic volatility. AIMS Mathematics, 2023, 8(10): 25164-25194. doi: 10.3934/math.20231284 |
[3] | Jiajia Zhao, Zuoliang Xu . Calibration of time-dependent volatility for European options under the fractional Vasicek model. AIMS Mathematics, 2022, 7(6): 11053-11069. doi: 10.3934/math.2022617 |
[4] | Din Prathumwan, Thipsuda Khonwai, Narisara Phoochalong, Inthira Chaiya, Kamonchat Trachoo . An improved approximate method for solving two-dimensional time-fractional-order Black-Scholes model: a finite difference approach. AIMS Mathematics, 2024, 9(7): 17205-17233. doi: 10.3934/math.2024836 |
[5] | Jafar Biazar, Fereshteh Goldoust . Multi-dimensional Legendre wavelets approach on the Black-Scholes and Heston Cox Ingersoll Ross equations. AIMS Mathematics, 2019, 4(4): 1046-1064. doi: 10.3934/math.2019.4.1046 |
[6] | Shoude Huang, Xin-Jiang He . An analytical approximation formula for European option prices under a liquidity-adjusted non-affine stochastic volatility model. AIMS Mathematics, 2022, 7(6): 10364-10377. doi: 10.3934/math.2022577 |
[7] | Teh Raihana Nazirah Roslan, Sharmila Karim, Siti Zulaiha Ibrahim, Ali Fareed Jameel, Zainor Ridzuan Yahya . Stochastic pricing formulation for hybrid equity warrants. AIMS Mathematics, 2022, 7(1): 398-424. doi: 10.3934/math.2022027 |
[8] | Shou-de Huang, Xin-Jiang He . Analytical approximation of European option prices under a new two-factor non-affine stochastic volatility model. AIMS Mathematics, 2023, 8(2): 4875-4891. doi: 10.3934/math.2023243 |
[9] | Yilihamujiang Yimamu, Zui-Cha Deng, Liu Yang . An inverse volatility problem in a degenerate parabolic equation in a bounded domain. AIMS Mathematics, 2022, 7(10): 19237-19266. doi: 10.3934/math.20221056 |
[10] | Hijaz Ahmad, Muhammad Nawaz Khan, Imtiaz Ahmad, Mohamed Omri, Maged F. Alotaibi . A meshless method for numerical solutions of linear and nonlinear time-fractional Black-Scholes models. AIMS Mathematics, 2023, 8(8): 19677-19698. doi: 10.3934/math.20231003 |
In this paper a finite difference method (FDM) is provided for pricing perpetual timer options under the Heston volatility model. Considering the degeneracy of the pricing equation, we first prove the existence and uniqueness of the solution of the pricing problem with a new notion of boundary conditions at degenerate boundary and the infinity. Then we discuss the choice of artificial boundary value conditions and obtain a prior estimate of the internal error caused by the boundary value error. This estimate helps to choose appropriate artificial boundary values and solution domain to reduce internal error of the numerical solution. Furthermore, We build a FDM with second-order convergence for the pricing problem. Finally, we implement our method and show the visualization results.
Timer option is a kind of exotic option whose payoff is the same as a vanilla option while the expiration date is not given in advance, but is determined by the cumulative variance of the underlying asset. When the cumulative variance reaches a preset threshold, the timer option automatically expires. A timer option is called a perpetual timer option if the option will not expire until the threshold is hit. As mentioned by Sawyer [1], timer options were first introduced for trading by Société Générale and Investment Banking since 2007. At that time, such an exotic option provided a more flexible risk management tool which allows market participants specifying the volatility of underlying asset by choosing a certain cumulative variance threshold. Li [2] demonstrated that a timer option can be cheaper than a vanilla option with the same strike price and the same expected investment horizon.
Long before timer options were traded in financial markets, Neuberger [3] proposed the "mileage" option with variant expiration date and Bick [4] discussed the replication of such option under a continuous-time model. As mentioned by Carr & Lee [5], these works can be seen as pioneer works of timer options. There are three major approaches to pricing timer options. The most commonly used approach is pricing through the Monte Carlo method. Li [2] first proposed a Black-Scholes-Merton type formula for pricing timer options and implemented a Monte Carlo simulation under the Heston model. They employed an Euler scheme on the Bessel process with predictor and corrector to improve accuracy. Bernard & Cui [6] successfully reduced the timer option pricing problem to a one-dimensional problem where only the dynamics of the volatility needs to be simulated. They implemented their method for Heston model and Hull White model. Cui [7] further improved the simulation method. The second approach is to find a closed-form approximation formula through asymptotic analysis technique. Li & Mercurio [8] developed a closed-form approximation for timer options under Heston model and the 3/2-model with small volatility of variance. They also provided an analysis framework for general stochastic volatility models. Furthermore, Li & Mercurio [9,10] discussed the asymptotic approximation for both perpetual and finite time timer options. In a recent work of Wang et al. [11], the author provided a closed-form approximation formula for timer options under a second-order multi-scale stochastic volatility model. Kwok & Zheng [12] introduced a numerical Fourier transform algorithm for finite maturity timer options under the 3/2-model. The third approach is through the path-integral method, which is borrowed from quantum field theory. By the path-integral method, Liang et al. [13] provided the multivariate integral expression of the timer option price under the Heston model and the 3/2-model. Cui et al. [14] used a stochastic time change technique to develop explicit formula for prices of timer options in the Heston model and the 3/2 model. Zhang et al. [15] developed a closed-form pricing formula for timer options under the Hull White model using this method.
The goal of this paper is to provide the fourth approach for pricing perpetual timer option under the Heston model, which is the finite difference method, theoretically and empirically. Different from the three approaches introduced before, especially the most widely used Monte Carlo method, the finite difference method is not a single-point calculation method. It can not only directly give the entire pricing surface, but its calculation results can also be more conveniently used for the calculation of the Greek value surface. In the Monte Carlo method, a huge amount of numerical simulation is required to obtain the entire pricing surface, and the simulation results cannot be directly used for the calculation of the Greek values. In addition, our method can give a priori theoretical error estimates, while the Monte Carlo method can only give confidence intervals for simulation results, which is a posteriori estimate based on relative errors. Compared with the finite difference method, although the calculation efficiency of the asymptotic approximation approach is higher, the accuracy of the approximate formula is limited by specific model parameters and cannot be flexibly adjusted according to actual needs. The closed-form formula obtained by the path-integral method can guarantee high calculation accuracy, but it is very complicated to implement and cannot be used to calculate Greeks. In fact, the finite difference method is one of the most commonly used methods for option pricing problems. It has the advantages of easy implementation, flexibility and high efficiency. However, due to the degenerate nature of the perpetual timer option pricing equation, the common finite difference method can hardly work. Li [2] illustrated a numerical example for pricing timer option under the Heston model using the ADI scheme. The numerical results showed that the convergence of this method is problematic. Since then, there is almost no relevant literature discussing the application of finite difference method to the timer option pricing problem.
Let's recall the pricing PDE of the perpetual timer call option under the Heston model from Li & Mercurio [9]:
{Lu(x,y,t)=0,(x,y,t)∈R×(0,∞)×(0,T],u(x,y,0)=max{ex−K,0},(x,y)∈R×(0,∞), | (1.1) |
where the variables x,y represent the logarithmic price and instantaneous volatility of the underlying asset respectively, the variable t represents the time to maturity of the option, and the constant K represents the strike price of the option. The operator L is defined as
Lu=ut−(12uxx+σ22uyy+ρσuxy+(ry−12)ux+κ(θy−1)uy−ryu). | (1.2) |
The coefficients defining L are all constants and satisfy following conditions:
r>0,κ>0,θ>0,σ>0,σ2<2κθ,−1<ρ<1, | (1.3) |
where r is the risk-free rate, θ is the long-term mean of volatility, κ indicating the speed of mean reversion, σ is the volatility of volatility, and ρ is the correlation coefficient between volatility dynamics and underlying price dynamics. For more details about the financial meaning of these coefficients, we recommend Heston [16] and Li & Mercurio [9].
From (1.2) we can see that the equation shows degeneracy along the boundary at {y=0} since the coefficients of ux,uy and u explode there. We also notice that no boundary conditions are specified along {y=0}. Such degeneracy indicates that in order to ensure the existence and uniqueness of the solution of (1.1) we must add suitable asymptotic growth conditions. Furthermore, in order to obtain a numerical solution, we also need to add appropriate boundary conditions at the degenerate boundary.
The main contribution of this paper are as follows. Firstly, we provide an asymptotic growth condition which theoretically guarantee the existence and uniqueness of the solution of the pricing problem (1.1), and we discuss the choice of artificial boundary value conditions at the degenerate boundary and infinity. We show that the error introduced by artificial boundary value conditions decays rapidly inside the solution domain under the control of the growth condition. We provide an a priori estimate of this error. Secondly, we developed a finite difference scheme with second-order convergence in spatial directions to solve the pricing problem numerically. The convergence of our scheme follows from a discrete Aleksandrov-Bakelman-Krylov maximum principle given by Kuo & Trudinger [17]. Finally, we conduct numerical experiments with our method.
The remainder of this paper is organized as follows. In section 2, we discuss the asymptotic condition and give a priori error estimation of the artificial boundary value error. In section 3, we provide the finite difference scheme and prove the convergence. In section 4 we conduct numerical experiments to show the convergence and accuracy of our method. Section 5 concludes the paper.
We first make some notations and state the asymptotic growth condition in the following subsection.
Let domain D⊂Rn, I=(a,b] is an interval on R1, Q=D×I, we thus define boundaries of Q as follows:
∂bQ=D×{a},∂xQ=∂D×I,∂pQ=∂bQ∪∂xQ. |
We also define the the closure and interior of Q respectively as:
ˉQ=Q∪∂pQ,Qo=ˉQ∖∂pQ. |
For positive parameters δ,R,R1,R2, we define domains that will be used in this article as follows.
Ω=R×(0,∞),Ωδ,R=[−R,R]×(δ,R],Ωδ,R1,R2=[−R1,R1]×(δ,R2]. | (2.1) |
For all subdomain D of Ω and constant T>0, we denote
(D)T=DT=D×(0,T]. |
For all real number a, we define
a+=max{a,0},a−=(−a)+. |
For any domain D⊂Ω×R, we can define a function class H(D):
H(D)={f:∃C>0 s.t. |f(x,y,t)|≤C(ex+1)}, | (2.2) |
where f(x,y,t):D↦R. We claim
|f|≤C(ex+1) | (2.3) |
as the asymptotic growth condition.
The asymptotic growth condition (2.3) actually provides a special boundary condition at the degenerate boundary. With such growth condition we can prove a comparison principle in ΩT, and the uniqueness follows.
Theorem 2.1 (Existence and Uniqueness). There exists unique solution u∈H(ˉΩT) of the pricing problem (1.1).
Remark 1. The existence for the solution of (1.1) can be obtained by directly using a classical Schauder estimate (here we refer to Lieberman [18]) and standard diagonal sub-sequence procedures. And the uniqueness follows directly from the comparison principle in the following Lemma 2.2.
Lemma 2.2 (Comparison Principle). If u,v∈H(ˉΩT) such that:
{Lu(x,y,t)−Lv(x,y,t)≤0,(x,y,t)∈ΩT,u(x,y,t)−v(x,y,t)≤0,(x,y,t)∈∂bΩT, |
then u≤v in ΩT.
Proof. By the definition of H in (2.2), we obtain that (u−v)∈H(ˉΩT), and thus there exists a constant C>0 such that
|u(x,y,t)−v(x,y,t)|≤C(ex+1),(x,y,t)∈ˉΩT. |
We define constants n0,λ as:
n0=|2rκθ|+1,λ=12(1−2κθσ2), | (2.4) |
and constants N1,N2,N3 as:
N1=(κ−ρσ)22σ2+1,N2=12n20σ2+(2σ+κ)n0+1,N3=2(κ−ρσ)2+2κθ+σ2+1. | (2.5) |
We also denote constant N=max(N1,N2,N3). To get the comparison principle, we define a barrier function p(x,y,t) as follows
p(x,y,t)=yλex+Nt+(e2x+e−x)eNt−n0y+(y2+1)ex+Nt. |
By directly calculation, we can check that P(x,y,t)>0 and LP(x,y,t)>0 in ΩT. Furthermore, we have:
lim(x,y,t)→∂xΩTex+1P(x,y,t)=0. | (2.6) |
Considering that u,v∈H(ˉΩT), for all ϵ>0 there exists δ0>0,R0>0 such that
u(x,y,t)−v(x,y,t)−ϵP(x,y,t)≤0,(x,y,t)∈∂p(Ωδ,R)T, |
as long as 0<δ<δ0 and R≥R0. We also know that
L(u(x,y,t)−v(x,y,t)−ϵP(x,y,t))≤0,(x,y,t)∈(Ωδ,R)oT. |
Then by the standard comparison principle for non-degenerate equations, we have
u(x,y,t)−v(x,y,t)≤ϵP(x,y,t),(x,y,t)∈(Ωδ,R)T. |
For all ϵ>0, let δ→0 and R→∞, we can see that
u(x,y,t)−v(x,y,t)≤ϵP(x,y,t),(x,y,t)∈ΩT. |
Then
u(x,y,t)−v(x,y,t)≤0, |
is achieved by letting ϵ→0. ■
Corollary 2.3. The solution u∈H(ˉΩT) of (1.1) satisfies the following condition
(ex−K)+≤u≤ex. | (2.7) |
Remark 2. Since Lex=0,L0=0 and LK=ry−1K>0, the condition (2.7) follows directly from Lemma 2.2. And this is also in line with financial intuition, that is, the price of the call option should not be less than its final payoff and should not exceed the underlying asset itself.
In practice, finding the numerical solution for the pricing problem (1.1) forces us to make a cut-off from the infinity of y=∞, x=±∞ and from the degenerate boundary of y=0. This cut-off leads to a bounded domain problem on (Ωδ,R1,R2)T. We have to provide additional artificial boundary conditions on ∂x(Ωδ,R1,R2)T. The most straightforward way is that we directly provide the exact solution of the pricing problem (1.1) on ∂x(Ωδ,R1,R2)T as the boundary value condition, however, we do not know the exact solution there.
In this section we prove that as long as the artificial boundary value satisfies the asymptotic growth condition (2.3), the error brought by the boundary value is convergent within the interior domain, which means we may only need to provide a not too outrageous boundary value to ensure sufficient accuracy inside the solution area.
We first setup the boundary error equation. The bounded domain pricing problem with artificial boundary value condition ψa(x,y,t) is shown as follows:
{Lub(x,y,t)=0,(x,y,t)∈(Ωδ,R1,R2)T,ub(x,y,t)=(ex−K)+,(x,y,t)∈∂b(Ωδ,R1,R2)T,ub(x,y,t)=ψa(x,y,t),(x,y,t)∈∂x(Ωδ,R1,R2)T. | (2.8) |
Let u be the solution of the original pricing problem (1.1). We define the error function as
eb(x,y,t)=u(x,y,t)−ub(x,y,t). |
Naturally, eb solves following problem:
{Leb(x,y,t)=0,(x,y,t)∈(Ωδ,R1,R2)T,eb(x,y,t)=0,(x,y,t)∈∂b(Ωδ,R1,R2)T,eb(x,y,t)=u(x,y,t)−ψa(x,y,t),(x,y,t)∈∂x(Ωδ,R1,R2)T. | (2.9) |
We provide a priori estimate for eb in the following Theorem 2.4.
Theorem 2.4 (Boundary Error Estimation). Let eb be the solution of (2.9), where the artificial boundary value ψa satisfies the asymptotic growth condition (2.3). We assume there is a constant C>0 such that |ψa|≤C(ex+1) in ∂p(Ωδ,R1,R2)T and choose constants λ,n0,N be the same as in Lemma 2.6. Then we have the following estimate:
|eb(x,y,t)|≤E0(x,y,t)+E1(x,y,t)+E2(x,y,t), | (2.10) |
where
E0(x,y,t)=δ−λ[(C+1)+CeR1]ex+Ntyλ,E1(x,y,t)=e−R1(4C+2)(e2x+e−x)eNt+n0R2,E2(x,y,t)=1R22+1(C+1)(y2+1)ex+Nt+e−R2n0Cey0n+Nt. | (2.11) |
Remark 3. E0, E1 and E2 control the boundary value errors from y=δ, x=±R1 and y=R2 respectively. In practice, for all error tolerance level ϵ>0 at an interior point (x,y,t) of (Ωδ,R1,R2)T, we can select sufficiently large R2, R1 and a sufficiently small δ accordingly to ensure |eb(x,y,t)|≤ϵ.
Corollary 2.5. Let eb be the solution of (2.9), if ψa satisfies the following condition:
(ex−K)+≤ψa(x,y,t)≤ex, |
then we have a better estimates as follows:
|eb(x,y,t)|≤[δ−λyλex+e−2R1(e2x+e−x)en0R2+ey0−R2n0]eNt. | (2.12) |
To show the proof of Theorem 2.4, we first discuss the boundary value error from different directions separately. Since both u and ψa satisfy the asymptotic growth condition (2.3), we discuss the constant-controlled boundary value error and the exponentially growing boundary value error in the following Lemma 2.6 and Lemma 2.7, respectively.
Lemma 2.6. If eb(x,y,t) is the solution of (2.9), and we choose constants λ,n0,N be the same as in (2.4) and (2.5), then we have following estimates:
(i) If |u(x,y,t)−ψa(x,y,t)|≤1∂p(Ωδ,R1,R2)T∩{y=δ} on ∂p(Ωδ,R1,R2)T, then we have
|eb(x,y,t)|≤δ−λeR1yλex+Nt. | (2.13) |
(ii) If |u(x,y,t)−ψa(x,y,t)|≤1∂p(Ωδ,R1,R2)T∩{x=−R1orx=R1} on ∂p(Ωδ,R1,R2)T, then we have
|eb(x,y,t)|≤e−R1+n0R2(e2x+e−x)eNt−n0y. | (2.14) |
(iii) If |u(x,y,t)−ψa(x,y,t)|≤1∂p(Ωδ,R1,R2)T∩{y=R2} on ∂p(Ωδ,R1,R2)T, then we have
|eb(x,y,t)|≤e(y−R2)/n0+Nt. | (2.15) |
Lemma 2.7. If the error function eb(x,y,t) is the solution of (2.9) and we choose constants λ,n0,N be the same as in Lemma 2.6, then we have following estimates:
(i) If |u(x,y,t)−ψa(x,y,t)|≤ex1∂p(Ωδ,R1,R2)T∩{y=δ} on ∂p(Ωδ,R1,R2)T, then we have:
|eb(x,y,t)|≤δ−λyλex+Nt. | (2.16) |
(ii) If |u(x,y,t)−ψa(x,y,t)|≤ex1∂p(Ωδ,R1,R2)T∩{x=R1} holds on ∂p(Ωδ,R1,R2)T, then we have:
|eb(x,y,t)|≤e−R1+n0R2(e2x+e−x)eNt−n0y. | (2.17) |
(iii) If |u(x,y,t)−ψa(x,y,t)|≤ex1∂p(Ωδ,R1,R2)T∩{x=−R1} holds on ∂p(Ωδ,R1,R2)T, then we have:
|eb(x,y,t)|≤e−2R1+n0R2(e2x+e−x)eNt−n0y. | (2.18) |
(iv) If |u(x,y,t)−ψa(x,y,t)|≤ex1∂p(Ωδ,R1,R2)T∩{y=R2} holds on ∂p(Ωδ,R1,R2)T, then we have:
|eb(x,y,t)|≤11+R22(1+y2)ex+Nt. | (2.19) |
Remark 4. The proof of Lemma 2.6 and Lemma 2.7 can be achieved by directly using a standard comparison principle for non-degenerate equation. We omit the details of the proof in this paper.
Proof of Theorem 2.4. From the Theorem 2.1 we know that the solution u(x,y,t) of the pricing problem (1.1) lies between (ex−K)+ and ex. Since |ψa|≤C(ex+1), we have
|u(x,y,t)−ψa(x,y,t)|≤|ψa(x,y,t)|+|u(x,y,t)|≤(C+1)ex+C, |
which indicates that the artificial boundary error is controlled by a linear combination of ex and 1 at each direction. Then we employ the boundary error estimates from Lemma 2.6 and Lemma 2.7 and find that:
|eb(x,y,t)|≤(C+1)δ−λyλex+Nt+Cδ−λeR1yλex+Nt+(2C+1)e−R1+n0R2(e2x+e−x)eNt−n0y+(2C+1)e−2R1+n0R2(e2x+e−x)eNt−n0y+C+1R22+1(y2+1)ex+Nt+Cey0−R2n0+Nt≤δ−λ[(C+1)+CeR1]ex+Ntyλ+e−R1(4C+2)(e2x+e−x)eNt+n0R2+1R22+1(C+1)(y2+1)ex+Nt+e−R2n0Cey0n+Nt=E0(x,y,t)+E1(x,y,t)+E2(x,y,t). |
■
Remark 5. Since u lies between (ex−K)+ and ex, it is reasonable to choose ψa accordingly, and thus
|ψa(x,y,t)−u(x,y,t)|≤ex−(ex−K)+≤min{ex,K}. |
Then Corollary 2.5 follows from a similar procedure as in the proof of Theorem 2.4.
In this section, we provide a finite difference scheme for solving the bounded domain problem (2.8). Our scheme is quite different from the ADI scheme, which perform poorly, as shown by Li [2]. We prove the convergence of our scheme through a discrete maximum principle given by Kuo & Trudinger [17].
We define meshing parameters as follows:
h>0,τ>0. |
For all i,j,n∈N, we set
x0=−R1,xi=xi−1+h,y0=δ,yj=yj−1+σh,t0=0,tn=tn−1+τ. | (3.1) |
We denote Eτ,h as a time-space mesh on (Ωδ,R1,R2)T, where Eτ,h⊂(Ωδ,R1,R2)T and consists of meshing points from (3.1), and we also denote u(xi,yj,tn) as unij. Then we define the following difference operators:
δtunij=1τ(uni,j−un−1i,j),δxunij=14h(uni+1,j+1−uni−1,j+1+uni+1,j−1−uni−1,j−1),δyunij=14σh(uni+1,j+1−uni+1,j−1+uni−1,j+1−uni−1,j−1),δxxunij=14h2(uni+1,j+1+uni−1,j+1+uni+1,j−1+uni−1,j−1)+12h2(uni−1,j+uni+1,j−uni,j+1−uni,j−1−2uni,j),δyyunij=14σ2h2(uni+1,j+1+uni−1,j+1+uni+1,j−1+uni−1,j−1)+12σ2h2(uni,j−1+uni,j+1−uni−1,j−uni+1,j−2uni,j),δxyunij=uni+1,j+1−uni+1,j−1−uni−1,j+1+uni−1,j−14σh2. | (3.2) |
The finite difference operator LΔ corresponding to L is defined as follows:
LΔunij=−δtunij+(12δxxunij+σ22δyyunij+ρσδxyunij)+(ry−12)δxunij+(κθy−κ)δyunij−ryunij. | (3.3) |
Using (3.2), we can write the definition (3.3) of LΔ as follows:
LΔuni,j=a1uni+1,j+1+a2uni−1,j−1+a3uni+1,j−1+a4uni−1,j+1+a5uni,j+a6un−1i,j | (3.4) |
where parameters a1−a6 are defined as:
a1=1+ρ4h2+(rσ+κθy−σ2−κ)14σh,a2=1+ρ4h2−(rσ+κθy−σ2−κ)14σh,a3=1−ρ4h2+(rσ−κθy−σ2+κ)14σh,a4=1−ρ4h2−(rσ−κθy−σ2+κ)14σh,a5=−(1τ+1h2+ry),a6=1τ. | (3.5) |
It can be seen that our finite difference scheme is actually a five-point scheme. Then we consider the finite difference problem corresponding to (2.8):
{LΔuτ,h(x,y,t)=0,(x,y,t)∈Eoτ,h,uτ,h(x,y,t)=(ex−K)+,(x,y,t)∈∂bEτ,h,uτ,h(x,y,t)=ψa(x,y,t),(x,y,t)∈∂xEτ,h. | (3.6) |
where uτ,h is a real functions defined on Eτ,h and ψa is the artificial boundary value condition. Here we present an algorithm for solving the finite difference problem (3.6).
Algorithm 1 Solving the finite difference equation |
Input: Geometric parameters R1,R2,δ,τ,h,T and equation coefficients r,κ,θ,σ,ρ. |
Output: uNij on Eτ,h, where τN=T. |
1: Meshing on (Ωδ,R1,R2)T according to formula (3.1) to obtain Eτ,h |
2: Set n=0 and initialize unij for each i,j with initial and boundary value conditions. |
3: for all n≤N do |
4: Update un+1ij for each interior point by solving the linear systems obtained from Equation (3.4) and (3.6). |
5: Update n=n+1 |
6: end for |
Considering the sparsity of the coefficient matrix, memory usage can be reduced by matrix compression techniques. Linear systems can be solved using the Jacobi iteration method.
Remark 6. If ub is the solution of (2.8) we can give the truncation error of LΔ by using Taylor expansion as:
|LΔuτ,h(x,y,t)−LΔub(x,y,t)|=O(τ+h2),(x,y,t)∈(Ωδ,R1,R2)T, | (3.7) |
which indicates that our scheme has second-order accuracy in the spatial directions.
Theorem 3.1. If uτ,h is the solution of the finite difference problem (3.6) and ub is the solution of bounded domain pricing problem (2.8) then uτ,h converge to ub uniformly on Eτ,h.
Theorem 3.2. If u(x,y,t) is the solution of (1.1), we choose two artificial boundary conditions ψ(u)a(x,y,t)=ex and ψ(l)a(x,y,t)=(ex−K)+. Denote u(u)b and u(l)b as the solution for bounded domain problem (2.8) under ψa=ψ(u)a and ψa=ψ(l)a respectively. We also denote u(u)τ,h and u(l)τ,h as solutions for the finite difference problem (3.6) similarly. Define
ˆuτ,h=12(u(u)τ,h+u(l)τ,h), |
then we have:
|u(x,y,t)−ˆuτ,h(x,y,t)|≤12|u(u)τ,h(x,y,t)−u(l)τ,h(x,y,t)|+O(τ+h2). | (3.8) |
For the proof of Theorem 3.1, we first introduce a discrete Aleksandrov-Bakelman-Krylov maximum principle, which mainly follows from Kuo & Trudinger [17], in the following lemma.
Lemma 3.3 (Discrete Maximum Principle). If uτ,h is a real function defined on Eτ,h, LΔ defined as (3.3), LΔuτ,h≥0 in Eoτ,h and uτ,h≤0 on ∂pEτ,h then there exists h∗>0, such that as long as 0<h≤h∗ we have:
maxEτ,huτ,h≤0. | (3.9) |
Proof. According to Kuo & Trudinger [17], the inequality (3.9) holds as long as the operator LΔ is evolving, monotone, time-wise non-degenerate, spatially non-degenerate and weak-positive. We will follow their definitions and show that these conditions can be satisfied if h is sufficiently small.
Recalling LΔ in (3.4) and parameters a1−a6 defined by (3.5). According to Kuo & Trudinger [17], LΔ is evolving since we do not use point with t>tn in (3.4). And it is time-wise non-degenerate since the sum of the parameters for all points with t<tn is positive, that is, a6>0.
Now we check the spatially non-degenerate condition. Denote
ξ=(ξ1,ξ2)∈Ω,h=(h,σh),h−=(h,−σh), |
by definition from [17], LΔ is spatially non-degenerate if there is a constant c>0 such that for all ξ∈Ω the following inequality holds:
(a1+a2)(h⋅ξ)2+(a3+a4)(h−⋅ξ)2≥c|ξ|2. |
Since −1<ρ<1, this can be achieved by directly calculation:
(a1+a2)(h⋅ξ)2+(a3+a4)(h−⋅ξ)=1+ρ2h2(ξ⋅h)2+1−ρ2h2(ξ⋅h−)2,=ξ21+σ2ξ22+2σρξ1ξ2,=(1−|ρ|)(ξ21+σ2ξ22)+|ρ|(ξ1+ρ|ρ|σξ2)2,≥(1−|ρ|)min(1,σ2)|ξ|2. |
Then we check the monotone and weak-positive conditions. According to Kuo & Trudinger [17], LΔ is monotone if a1,a2,a3,a4 are all strictly positive, and LΔ is weak-positive if it is monotone and ∑5i=1ai<0. Noting that r,σ,κ,θ>0, |ρ|<1, σ2<2κθ and y>δ, we have:
ai≥1−|ρ|4h2−14σh(2rσ+σ22δ+σ2+κ),i∈{1,2,3,4}. |
It is easy to check ai>0 as long as
h<2δ(1−|ρ|)σσ2+2rσ+δ(σ+2κ), | (3.10) |
which indicates that LΔ is monotone. And is it quit directly to show that LΔ is weak-positive since ∑5i=1ai=−1τ−ry<0.
Finally, we define
h∗=2δ(1−|ρ|)σσ2+2rσ+δ(σ+2κ) |
to finish the proof. ■
Equipped with the Lemma 3.3, we can prove the convergence of our finite difference scheme.
Proof of Theorem 3.1. Define a perturbation function with parameter ϵ>0 to ub:
u(ϵ)b(x,y,t)=ub(x,y,t)+ϵt,(x,y,t)∈Eτ,h, |
We can check that
LΔu(ϵ)b−LΔuτ,h=LΔub−LΔuτ,h+ϵLΔt≤LΔub−LΔuτ,h−ϵ(1+ryt)≤LΔub−LΔuτ,h−ϵ. |
Considering the truncation error given by (3.7), for all ϵ>0, there exists τ∗,h∗ such that for all 0<τ<τ∗ and 0<h<h∗, we have
LΔub−LΔuτ,h−ϵ≤0, |
and thus
LΔu(ϵ)b−LΔuτ,h≤0, |
Since u(ϵ)b(x,y,t)=ub(x,y,t)+ϵt≥uτ,h at the boundary ∂pEτ,h, by Lemma 3.3, we have:
uτ,h(x,y,t)−u(ϵ)b(x,y,t)≤0,(x,y,t)∈Eτ,h. | (3.11) |
Symmetrically, we define u(−ϵ)b(x,y,t)=ub(x,y,t)+ϵt. One can check that
uτ,h(x,y,t)−u(−ϵ)b(x,y,t)≥0,(x,y,t)∈Eτ,h. | (3.12) |
Then we can say for all ϵ>0, there exists τ∗,h∗ such that as long as 0<τ<τ∗ and 0<h<h∗,
|uτ,h(x,y,t)−ub(x,y,t)|≤ϵt≤ϵT,(x,y,t)∈Eτ,h, |
which implies the convergence. ■
Proof of Theorem 3.2. According to Theorem 2.1, we know that
(ex−K)+≤u(x,y,t)≤ex. |
Then by the standard comparison principle, we can check that
u(l)b(x,y,t)≤u(x,y,t)≤u(u)b(x,y,t),(x,y,t)∈(Ωδ,R1,R2)T. |
Together with (3.7) and the Theorem 3.1, we have:
|u(x,y,t)−ˆuτ,h(x,y,t)|=12|(u−u(u)τ,h)+(u−u(l)τ,h)|=12|(u−u(u)b+u(u)b−u(u)τ,h)+(u−u(l)b+u(l)b−u(l)τ,h)|=12|(u−u(u)b)+(u−u(l)b)|+12|(u(u)b−u(u)τ,h)+(u(l)b−u(l)τ,h)|=12|(u−u(u)b)+(u−u(l)b)|+O(τ+h2)≤12|u(u)b−u(l)b|+O(τ+h2). |
■
In this part we implement some numerical experiments to verify the results established by the previous sections. We choose the parameters of L from Liang et al. [13], shown in Table 1.
θ | κ | σ | ρ | r | T |
0.09 | 2 | 0.375 | −0.5 | 0.015 | 0.087 |
We will solve the finite difference problem (3.6), which is corresponding to the bounded domain problem (2.8) and the original pricing problem (1.1), in two cases, K=0 and K=1, respectively. Then we show the effect of artificial boundary value error and the accuracy and convergence speed of the numerical method.
We start from a trivial case, in which the strike price K=0. With K=0, one can check that u(x,y,t)=ex is the solution to (1.1). Consider the following finite difference problem (3.6) corresponding to K=0:
{LΔuτ,h(x,y,t)=0,(x,y,t)∈Eoτ,huτ,h(x,y,t)=ex,(x,y,t)∈∂bEτ,h,uτ,h(x,y,t)=ψa(x,y,t),(x,y,t)∈∂xEτ,h, | (4.1) |
where Eτ,h is the time-space mesh of (Ωδ,R1,R2)T defined through (3.1).
Clearly, if we choose ψa(x,y,t)=ex, that is there is no error in the artificial boundary value, then the error between uτ,h and the exact solution u of (1.1) only comes from the truncation error, which is controlled by the meshing parameters τ and h. In this part, we first set ψa=ex and calculate uτ,h with different τ,h to verify the convergence speed of our method. Then we fix τ and h, but deliberately choose ψa with error to observe the effect of boundary value errors.
Given τ0>0,h0>0, the spatial discrete error and time discrete error are defined as:
eh=max{|uτ0,h(x,y,T)−uτ0,h0(x,y,T)|:(x,y,T)∈Eτ0,h},eτ=max{|uτ,h0(x,y,T)−uτ0,h0(x,y,T)|:(x,y,T)∈Eτ,h0}, | (4.2) |
where τ>τ0,h>h0.
We implement our numerical method on Eτ,h⊂(Ωδ,R1,R2)T with geometric parameters δ=0.01, R1=2,R2=1.
To show the convergence speed of eτ and eh, we choose ψa(x,y,t)=ex and set τ0=1.25×10−3, h0=5×10−2, and let τ varies from 1.45×10−3 to 2×10−3, h varies from 5×10−2 to 5×10−1. In particular, we choose τ,h to be integer multiples of τ0,h0 to ensure the grid points of Eτ,h are included in that of Eτ0,h0. The numerical results are shown in Figure 1. It can be seen that eτ and eh converges linearly with respect to τ and h2, respectively, which is consistent with the truncation error of the finite difference scheme given by (3.7).
Now we discuss the effect of boundary errors. The grid parameters τ0,h0 and the geometric constants δ,R1,R2 are set as before. Let u0(x,y,t) be the solution of (4.1) under ψa=ex on Eτ0,h0. Then we choose two artificial boundary value conditions ψ(1)a and ψ(2)a with error, where ψ(1)a=ex+1 has unit error and ψ(2)a=2ex has exponential error. We calculate uτ0,h0−u0 under the artificial boundary values ψ(1)a and ψ(2)a respectively. The numerical results are shown in Figure 2. We represent the values of uτ0,h0−u0 with a color gradient.
As can be seen from Figure 2(a), there is a constant error equal to one on the boundary of Eτ0,h0, but the error rapidly decays to a level close to zero inside the solution domain far from the boundary. Figure 2(b) shows that the exponential error have a similar decay. These results are consistent with the conclusion in subsection 2.3 that the boundary value errors converge rapidly inside the solution domain.
In this part, we consider a non-trivial case with K=1. The corresponding finite difference problem is defined as follows:
{LΔuτ1,h1(x,y,t)=0,(x,y,t)∈Eoτ1,h1uτ1,h1(x,y,t)=(ex−1)+,(x,y,t)∈∂bEτ1,h1,uτ1,h1(x,y,t)=ψa(x,y,t),(x,y,t)∈∂xEτ1,h1, | (4.3) |
where Eτ1,h1 is the time-space mesh of (Ωδ,R1,R2)T. The grid parameters τ1=2×10−5,h1=2.5×10−3 and the geometric parameters δ=2.5×10−4,R1=2, R2=1. We follow the statement of the Theorem 3.2 and choose two artificial boundary value conditions ψ(u)a=ex, ψ(l)a=(ex−1)+, then u(u)τ1,h1,u(l)τ1,h1 and ˆuτ1,h1 are defined accordingly. Figure 3 shows the difference between u(u)τ1,h1 and u(l)τ1,h1. The color gradient represent values of log(|u(u)τ1,h1(x,y,T)−u(l)τ1,h1(x,y,T)|).
The contour lines show the difference between the solutions under two different boundary value functions gradually diminish when leaving the boundary. Typically, such difference is less than 1×10−5 in the area enclosed by the innermost contour, where we can find that the error comes from boundary values is less than 5×10−6 and it has the same magnitude as the truncation error. In practical application, such small error is negligible.
To further demonstrate the accuracy of our method, we compare our results with those obtained from the Monte Carlo method, the asymptotic approximation method, and the path-integral method, respectively. Results obtained from the Monte Carlo method and the path-integral method are given by Liang et al. [13], and the results obtained from the asymptotic approximation method are given by Li and Mecurio [9]. In their work, numerical solutions for the pricing problem (1.1) are provided with the same parameters as ours at one single point: u(x0,y0,T)=0.125789, where x0=0 and y0=0.087. Since y=0.087 is not on the mesh grid of Eτ1,h1, we find the numerical solution there by a standard second order interpolation and the results are shown in Figure 4.
As can be seen from Figure 4(a), the red line representing u(l)τ1,h1 and the green line representing u(u)τ1,h1 are almost indistinguishable from the blue line representing ˆuτ1,h1 when x∈[−1,1]. Figure 4(b) further shows the difference between u(u)τ1,h1 and u(l)τ1,h1, which according to (3.8) mainly controls the accuracy of ˆuτ1,h1. Typically, at the point of (x0,y0,T), the results of each method are compared in Table 2 below. Since the path-integral method is considered to be a relatively accurate method by Li and Mercurio [9], we can use its results as a benchmark to calculate the relative errors of the results of the other three methods.
Path-integral Method | Monte Carol Method | Asymptotic Approximation Method | Out Finite Difference Method | |
Numerical Results | 0.125789 | 0.125668 | 0.125815 | 0.125786 |
Relative Errors | 0.000% | −0.0960% | 0.0210% | −0.0023% |
It can be seen that the relative error of our method is an order of magnitude smaller than the Monte Carlo method and the asymptotic approximation method. The gap between our result and the results of the path-integral method proposed by Liang et al. [13] is only 3×10−6, which is negligible in practical applications. Therefore our method can be considered to have sufficient accuracy.
It is worth noting that, unlike the path-integral method proposed by Liang et al. or the asymptotic approximation method proposed by Li and Mercurio, our method does not give the solution at a single point, but the solution at every point on Eτ1,h1. If we choose a bounded domain D=[−0.5,0.5]×[0.05,0.5], which is enclosed by the innermost contour of Figure 3. The option prices are shown in Figure 5. The error in D is on the order of 10−5.
In this paper, we provide a finite difference method for pricing perpetual timer options under Heston volatility model. Since the pricing equation is degenerate, we first prove that the solution of the pricing problem exists and is unique under the asymptotic growth condition. Then we give a priori estimate of the artificial boundary value error for the bounded domain problem (2.8) and show that the solution of (2.8) internally converges to the solution to the original pricing problem (1.1). Furthermore, we provide a finite difference scheme with second-order convergence to solve the bounded domain problem (2.8). We implement our finite difference method with parameters in Liang et al. [13], and visualize the convergence and accuracy of this method.
The authors declare no conflict of interest.
[1] | N. Sawyer, SG CIB launches timer options, Risk, 20 (2007), 6–7. |
[2] | Managing volatility risk, C. Li., Doctoral Dissertation of Columbia University, 2010. Available From: http://www.math.columbia.edu/thaddeus/theses/2010/li.pdf |
[3] | A. Board, Stochastic modelling and applied probability, Berlin: Springer, 2005. https://doi.org/10.1007/978-0-387-22593-7_1 |
[4] |
A. Bick, Quadratic-variation-based dynamic strategies, Manage. Sci., 41 (1995), 722–732. https://doi.org/10.1287/mnsc.41.4.722 doi: 10.1287/mnsc.41.4.722
![]() |
[5] |
P. Carr, R. Lee, Volatility derivativess, Annual Review of Financial Economics, 1 (2009), 319–339. https://doi.org/10.1146/annurev.financial.050808.114304 doi: 10.1146/annurev.financial.050808.114304
![]() |
[6] |
C. Bernard, Z. Cui, Pricing timer options, J. Comput. Financ., 15 (2011), 69–104. https://doi.org/10.21314/JCF.2011.228 doi: 10.21314/JCF.2011.228
![]() |
[7] |
C. Bernard, Z. Cui, D. McLeish, On the martingale property in stochastic volatility models based on time-homogeneous diffusions, Math. Financ., 27 (2014), 194–223. http://dx.doi.org/10.1111/mafi.12084 doi: 10.1111/mafi.12084
![]() |
[8] | M. Li, F. Mercurio, Time for a timer, Risk, 26 (2013), 76–81. |
[9] |
M. Li, F. Mercurio, Closed-form approximation of perpetual timer option prices, International Journal Of Theoretical And Applied Finance, 17 (2014), 1450026. https://doi.org/10.1142/s0219024914500265 doi: 10.1142/s0219024914500265
![]() |
[10] |
M. Li, F. Mercurio, Analytic approximation of finite-maturity timer option prices, J. Futures Markets, 35 (2015), 245–273. https://doi.org/10.1002/fut.21659 doi: 10.1002/fut.21659
![]() |
[11] |
X. Wang, S. Wu, X. Yue, Pricing timer options: second-order multiscale stochastic volatility asymptotics, ANZIAM J., 63 (2021), 249–267. https://doi.org/10.21914/anziamj.v63.15291 doi: 10.21914/anziamj.v63.15291
![]() |
[12] |
Y. Kwok, W. Zheng, Timer Options, Pricing Models of Volatility Products Exotic Variance Derivatives, 16 (2022), 217–248. https://doi.org/10.1201/9781003263524-6 doi: 10.1201/9781003263524-6
![]() |
[13] |
L. Liang, D. Lemmens, J. Tempere, Path integral approach to the pricing of timer options with the Duru-Kleinert time transformation, Phys. Rev. E, 83 (2011), 056112. https://doi.org/10.1103/physreve.83.056112 doi: 10.1103/physreve.83.056112
![]() |
[14] |
Z. Cui, J. L. Kirkby, G. Lian, D. Nguyen, Integral representation of probability density of stochastic volatility models and timer options, Int. J. Theor. Appl. Fin., 20 (2017), 1750055. https://doi.org/10.1142/s0219024917500558 doi: 10.1142/s0219024917500558
![]() |
[15] |
J. Zhang, X. Lu, Y. Han, Pricing perpetual timer option under the stochastic volatility model of Hull–White, ANZIAM J., 58 (2017), 406–416. https://doi.org/10.1017/s1446181117000177 doi: 10.1017/s1446181117000177
![]() |
[16] |
S. Heston, A closed-form solution for options with stochastic volatility with applications to bond and currency options, Rev. Financ. Stud., 6 (1993), 327–343. https://doi.org/10.1093/rfs/6.2.327 doi: 10.1093/rfs/6.2.327
![]() |
[17] |
H. Kuo, N. Trudinger, On the Krylov maximum principle for discrete parabolic schemes, Tamkang J. Math., 40 (2009), 437–450. https://doi.org/10.5556/j.tkjm.40.2009.607 doi: 10.5556/j.tkjm.40.2009.607
![]() |
[18] | G. M. Lieberman, Second order parabolic differential equations, New York: World scientific, 1996. https://doi.org/10.1142/3302 |
1. | Mijin Ha, Donghyun Kim, Ji-Hun Yoon, Pricing of timer volatility-barrier options under Heston’s stochastic volatility model, 2025, 457, 03770427, 116310, 10.1016/j.cam.2024.116310 |
θ | κ | σ | ρ | r | T |
0.09 | 2 | 0.375 | −0.5 | 0.015 | 0.087 |
Path-integral Method | Monte Carol Method | Asymptotic Approximation Method | Out Finite Difference Method | |
Numerical Results | 0.125789 | 0.125668 | 0.125815 | 0.125786 |
Relative Errors | 0.000% | −0.0960% | 0.0210% | −0.0023% |