This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Unbiased estimators for the Heston model with stochastic interest rates

Chao Zheng Email: chao.zheng12@gmail.com. School of Data Sciences, Zhejiang University of Finance and Economics, Hangzhou, China Jiangtao Pan Email: 211510011019@zufe.edu.cn School of Data Sciences, Zhejiang University of Finance and Economics, Hangzhou, China
Abstract

We combine the unbiased estimators in Rhee and Glynn (Operations Research: 63(5), 1026-1043, 2015) and the Heston model with stochastic interest rates. Specifically, we first develop a semi-exact log-Euler scheme for the Heston model with stochastic interest rates. Then, under mild assumptions, we show that the convergence rate in the L2L^{2} norm is O​(h)O(h), where hh is the step size. The result applies to a large class of models, such as the Heston-Hull-While model, the Heston-CIR model and the Heston-Black-Karasinski model. Numerical experiments support our theoretical convergence rate.

Keywords: Heston model, stochastic interest rates, unbiased estimators, convergence rate, Heston-Hull-While model, Heston-CIR model

AMS subject classifications (2000): 60H35, 65C30, 91G60

1 Introduction

The classical Heston model (Heston [13]) is one of the fundamental models in finance, and it has been widely applied in various financial markets, such as the equity, fixed income and foreign exchange markets, due to its tractability in modelling the term structure of implied volatility. However, in this model, the interest rate is constant, and this assumption is often not appropriate for long-maturity options, because the long-term behaviour of the interest rate is typically far from constant. This phenomenon has been empirically investigated by Bakshi, Cao and Chen [2]. A natural extension of the Heston model is to incorporate a stochastic interest rate, which is usually referred to as the Heston model with stochastic interest rates. There are several contributions in this direction, such as Grzelak and Oosterlee [12], Van Haastrecht and Pelsser [21] and references therein.

Under the Heston model with stochastic interest rates, the price of an option can be written as

𝔼​(eβˆ’βˆ«0Trt​dt​P​(S))\mathbb{E}\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S)\right)

where SS is the solution to the Heston model with stochastic interest rates and rtr_{t} is the underlying interest rate. Here, PP denotes the payoff functional of an option. We are interested in calculating this expectation, as for the majority of options, there are no closed-form formulas. A classical approach is to use a Monte Carlo method associated with a time-discrete scheme on SS for an approximate value. Recently, Rhee and Glynn [20] proposed several unbiased Monte Carlo estimators based on a randomization idea for a stochastic differential equation, which are unbiased versions of multilevel Monte Carlo estimators by Giles [9]. These unbiased estimators have a clear advantage over the standard Monte Carlo estimator, as the latter is typically biased when SS has to be approximated through a time-discrete scheme. To combine Rhee and Glynn’s estimators and the Heston model with stochastic interest rates, it is essential to develop a numerical scheme with a good convergence rate in the L2L^{2} norm. However, standard theorems, such as those in Kloeden and Platen [16], require that the drift and diffusion coefficients satisfy the global Lipschitz and linear growth assumptions, which are not satisfied by the Heston model with stochastic interest rates. Research on developing time-discrete schemes for the Heston model with stochastic interest rates is scarce. Cozma, Mariapragassam and Reisinger [6] proposed a different log-Euler scheme for the stochastic-local volatility model with stochastic rates and they demonstrated strong convergence without providing a rate. This is the only reference we are aware of concerning the convergence of Monte Carlo algorithms for the Heston model with stochastic interest rates.

In this article, we develop a semi-exact log-Euler scheme for the Heston model with stochastic interest rates, where the driven Brownian motion for interest rate models is independent of the driven Brownian motions for the Heston component. The scheme is an extension of those in Mickel and Neuenkirch [17] and Zheng [22] [23] for the classical Heston model. Under mild assumptions on interest rate models, we show that the underlying scheme converges with order one in the L2L^{2} norm. The types of options we consider include those with bounded and Lipschitz continuous payoffs and the digital option with a discontinuous payoff. There are two advantages of the scheme we develop. One is that the convergence rate is free of Feller’s index (except for the digital option), i.e., the convergence rate is valid for the full range of parameters in the Heston component of the Heston model with stochastic interest rates. The other advantage is that the convergence rate is higher than the usual convergence rate (one-half in the L2L^{2} norm) of the standard Euler scheme under standard assumptions. When a numerical scheme has a convergence rate higher than one-half, it is convenient to combine it into Rhee and Glynn’s unbiased estimators. Our result applies to a large class of models, including the Heston-Hull-While model, the Heston-CIR model and the Heston-Black-Karasinski model, among which the Heston-Hull-White model and the Heston-CIR model are particularly attractive in practical applications. We refer readers to Grzelak and Oosterlee [12] for further discussion.

The remainder of this article is organized as follows. In section 2, we review the Heston model with stochastic interest rates and develop a log-Euler scheme for it. Section 3 reviews the unbiased estimators from Rhee and Glynn [20]. In section 4, we derive the relevant convergence rate under several mild assumptions. Section 5 illustrates numerical results to support our theoretical analysis.

2 Heston model with stochastic interest rates

Let (Ξ©,β„±,(β„±t)tβ‰₯0,β„™)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\geq 0},\mathbb{P}) be a filtered probability space satisfying the usual assumptions. The Heston model with stochastic interest rates is of the form

d​St\displaystyle\mathrm{d}S_{t} =rt​St​d​t+Vt​St​(ρ​d​Wt1+1βˆ’Ο2​d​Wt2)\displaystyle=r_{t}S_{t}\mathrm{d}t+\sqrt{V_{t}}S_{t}(\rho\mathrm{d}W_{t}^{1}+\sqrt{1-\rho^{2}}\mathrm{d}W_{t}^{2})
d​Vt\displaystyle\mathrm{d}V_{t} =k​(ΞΈβˆ’Vt)​d​t+σ​Vt​d​Wt1,\displaystyle=k(\theta-V_{t})\mathrm{d}t+\sigma\sqrt{V_{t}}\mathrm{d}W_{t}^{1},

where (Wt1)tβ‰₯0(W^{1}_{t})_{t\geq 0} and (Wt2)tβ‰₯0(W^{2}_{t})_{t\geq 0} are two independent β„±t\mathcal{F}_{t}-adapted Brownian motions and the parameters k,ΞΈ,Οƒ>0k,\theta,\sigma>0 and ρ∈[βˆ’1,1]\rho\in[-1,1]. The initial values S0,V0>0S_{0},V_{0}>0. Here, (rt)tβ‰₯0(r_{t})_{t\geq 0} is a stochastic interest rate. If we replace rtr_{t} in the model with a constant, then it becomes the classical Heston model. The classical interest rate models and their generalizations can be found in Brigo and Mercurio [4]. Among them, a large class of interest rate models can be written as

d​rt=μ​(t,rt)​d​t+ϕ​(t,rt)​d​Wt3,\mathrm{d}r_{t}=\mu(t,r_{t})\mathrm{d}t+\phi(t,r_{t})\mathrm{d}W_{t}^{3},

where ΞΌ,Ο•:[0,T]×ℝ→ℝ\mu,\phi:[0,T]\times\mathbb{R}\rightarrow\mathbb{R} are continuous functions and (Wt3)tβ‰₯0(W_{t}^{3})_{t\geq 0} is a β„±t\mathcal{F}_{t}-adapted Brownian motion. We assume that (Wt3)tβ‰₯0(W_{t}^{3})_{t\geq 0} is independent of (Wt1)tβ‰₯0(W_{t}^{1})_{t\geq 0} and (Wt2)tβ‰₯0(W_{t}^{2})_{t\geq 0}. Furthermore, we assume that there is a unique solution to the equation of (rt)tβ‰₯0(r_{t})_{t\geq 0} above.

Let Xt=ln⁑(St)X_{t}=\ln(S_{t}). By using Ito^\hat{o}’s lemma, we have

d​Xt=(rtβˆ’12​Vt)​d​t+Vt​(ρ​d​Wt1+1βˆ’Ο2​d​Wt2).\mathrm{d}X_{t}=\left(r_{t}-\frac{1}{2}V_{t}\right)\mathrm{d}t+\sqrt{V_{t}}\left(\rho\mathrm{d}W_{t}^{1}+\sqrt{1-\rho^{2}}\mathrm{d}W_{t}^{2}\right).

Then substituting the equation of VtV_{t} into the equation above, we obtain

d​Xt=((rtβˆ’k​ρ​θσ)+(kβ€‹ΟΟƒβˆ’12)​Vt)​d​t+ρσ​d​Vt+1βˆ’Ο2​Vt​d​Wt2.\mathrm{d}X_{t}=\left(\left(r_{t}-\frac{k\rho\theta}{\sigma}\right)+\left(\frac{k\rho}{\sigma}-\frac{1}{2}\right)V_{t}\right)\mathrm{d}t+\frac{\rho}{\sigma}\mathrm{d}V_{t}+\sqrt{1-\rho^{2}}\sqrt{V_{t}}\mathrm{d}W_{t}^{2}.

Since (Vt)tβ‰₯0(V_{t})_{t\geq 0} is independent of (Wt2)tβ‰₯0(W^{2}_{t})_{t\geq 0}, the stochastic integral ∫0TVt​dWt2\int_{0}^{T}\sqrt{V_{t}}\mathrm{d}W_{t}^{2} is normally distributed with mean 0 and variance ∫0TVt​dt\int_{0}^{T}V_{t}\mathrm{d}t. Therefore, the solution at any finite time horizon T>0T>0 can be written as

XT\displaystyle X_{T} =X0+[∫0Trtdt+(ρ​kΟƒβˆ’12)∫0TVtdt+ρσ(VTβˆ’V0βˆ’kΞΈT)\displaystyle=X_{0}+\left[\int_{0}^{T}r_{t}\mathrm{d}t+\left(\frac{\rho k}{\sigma}-\frac{1}{2}\right)\int_{0}^{T}V_{t}\mathrm{d}t+\frac{\rho}{\sigma}\left(V_{T}-V_{0}-k\theta T\right)\right.
+1βˆ’Ο2∫0TVt​dtN]\displaystyle\quad\left.+\sqrt{1-\rho^{2}}\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}N\right] (1)

where NN is a standard normal random variable, that is independent of (Vt)t∈[0,T](V_{t})_{t\in[0,T]}. Note that NN is also independent of (rt)t∈[0,T](r_{t})_{t\in[0,T]}, because the driving Brownian motion (Wt3)t∈[0,T](W^{3}_{t})_{t\in[0,T]} of (rt)t∈[0,T](r_{t})_{t\in[0,T]} is independent of (Wt2)t∈[0,T](W^{2}_{t})_{t\in[0,T]} and (Vt2)t∈[0,T](V^{2}_{t})_{t\in[0,T]}. There are several integrals in equation (1) to be approximated.

It is known that VtV_{t} follows a scaled noncentral chi-squared distribution given VuV_{u} for any u∈[0,t)u\in[0,t), i.e.,

Vt​=d​σ2​(1βˆ’eβˆ’k​(tβˆ’u))4​k​χd2​(4​k​eβˆ’k​(tβˆ’u)Οƒ2​(1βˆ’eβˆ’k​(tβˆ’u))​Vu),V_{t}\overset{\mathrm{d}}{=}\frac{\sigma^{2}(1-\mathrm{e}^{-k(t-u)})}{4k}\chi_{d}^{2}\left(\frac{4k\mathrm{e}^{-k(t-u)}}{\sigma^{2}(1-\mathrm{e}^{-k(t-u)})}V_{u}\right),

where Ο‡d2​(Ξ»)\chi_{d}^{2}(\lambda) denotes a noncentral chi-squared random variable with degrees of freedom d=4​k​θσ2>0d=\frac{4k\theta}{\sigma^{2}}>0 and noncentrality parameter Ξ»>0\lambda>0 (see Glasserman [11]). Hence, VtV_{t} can be sampled exactly. Let (r^i​h)i=1,..,T/h(\hat{r}_{ih})_{i=1,..,T/h} be an approximate path of (rt)t∈[0,T](r_{t})_{t\in[0,T]}. It is convenient to approximate

∫0Trt​dtβ‰ˆβˆ‘i=0T/hβˆ’1r^i​h​h,∫0TVt​dtβ‰ˆβˆ‘i=0T/hβˆ’1Vi​h​h,\int_{0}^{T}r_{t}\mathrm{d}t\approx\sum_{i=0}^{T/h-1}\hat{r}_{ih}h,\quad\int_{0}^{T}V_{t}\mathrm{d}t\approx\sum_{i=0}^{T/h-1}V_{ih}h,

by using the Euler scheme based on step size hh. We denote by X^Th\hat{X}_{T}^{h} the approximated solution of XTX_{T}. Let S^Th:=eX^Th\hat{S}_{T}^{h}:=e^{\hat{X}_{T}^{h}}, and then S^Th\hat{S}_{T}^{h} is an approximation of STS_{T}.

3 Unbiased estimators for SDEs

In this section, we review the unbiased estimators introduced in Rhee and Glynn [20]. The prices of many options can be expressed as

𝔼​(Y):=𝔼​(eβˆ’βˆ«0Trt​dt​P​(ST))\mathbb{E}(Y):=\mathbb{E}\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})\right)

where PP is the payoff function and Y∈L2Y\in L^{2} (i.e., 𝔼​(Y2)<∞\mathbb{E}(Y^{2})<\infty). To estimate this expectation, Rhee and Glynn [20] proposed an estimator

Z=βˆ‘n=0𝒩Δnℙ​(𝒩β‰₯n)Z=\sum_{n=0}^{\mathcal{N}}\frac{\Delta_{n}}{\mathbb{P}(\mathcal{N}\geq n)}

where Ξ”n=Ynβˆ’Ynβˆ’1\Delta_{n}=Y_{n}-Y_{n-1} and YnY_{n}, nβˆˆβ„•n\in\mathbb{N}, approximates YY with step size T/2nT/2^{n} and Yβˆ’1=0Y_{-1}=0. In this article, we let

Yn=eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h​P​(S^Th),h=T/2n.Y_{n}=e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}P(\hat{S}_{T}^{h}),\quad h=T/2^{n}.

Here, 𝒩\mathcal{N} is a nonnegative integer-valued random variable that is independent of YnY_{n}. This estimator is usually referred to as the coupled sum estimator. Other unbiased estimators can be constructed in a similar way (see Rhee and Glynn [20]).

Suppose that YnY_{n} converges to YY in the L2L^{2} norm as nβ†’βˆžn\rightarrow\infty. Theorem 1 of Rhee and Glynn [20] showed that if

βˆ‘n=1βˆžπ”Όβ€‹[(Ynβˆ’1βˆ’Y)2]ℙ​(𝒩β‰₯n)<∞\sum_{n=1}^{\infty}\frac{\mathbb{E}\left[(Y_{n-1}-Y)^{2}\right]}{\mathbb{P}(\mathcal{N}\geq n)}<\infty (2)

then ZZ is an unbiased estimator of 𝔼​(Y)\mathbb{E}(Y) (i.e., 𝔼​(Z)=𝔼​(Y)\mathbb{E}(Z)=\mathbb{E}(Y)) with a finite variance. Furthermore, the average computational time of ZZ is proportional to

βˆ‘n=0∞2n​ℙ​(𝒩β‰₯n).\sum_{n=0}^{\infty}2^{n}\mathbb{P}(\mathcal{N}\geq n). (3)

Therefore, if 𝔼​[(Ynβˆ’Y)2]=O​(2βˆ’2​n​p)\mathbb{E}\left[(Y_{n}-Y)^{2}\right]=O(2^{-2np}) with p>1/2p>1/2 (here, pp is the convergence rate in the L2L^{2} norm), we can easily construct a distribution for 𝒩\mathcal{N} such that ℙ​(𝒩β‰₯n)=O​(2βˆ’n​(p+1/2))\mathbb{P}(\mathcal{N}\geq n)=O(2^{-n(p+1/2)}) to ensure that (2) and (3) are finite. The optimal distribution of 𝒩\mathcal{N} can be calculated by minimizing the product of the variance and the average computational time of ZZ; see Rhee and Glynn [20] and Zheng, Pan and Wang [24] for further discussion. Hence, it is important to investigate the convergence rate of 𝔼​[(Ynβˆ’Y)2]\mathbb{E}\left[(Y_{n}-Y)^{2}\right].

4 Convergence analysis

Recall that the interest rate (rt)t∈[0,T](r_{t})_{t\in[0,T]} follows the stochastic differential equation

d​rt=μ​(t,rt)​d​t+ϕ​(t,rt)​d​Wt3,\mathrm{d}r_{t}=\mu(t,r_{t})\mathrm{d}t+\phi(t,r_{t})\mathrm{d}W_{t}^{3},

where (Wt3)t∈[0,T](W_{t}^{3})_{t\in[0,T]} is independent of (Wt1)t∈[0,T](W_{t}^{1})_{t\in[0,T]} and (Wt2)t∈[0,T](W_{t}^{2})_{t\in[0,T]}. Let cc and cqc_{q} be constants regardless of their values, where cqc_{q} relies on qβ‰₯1q\geq 1. Our analysis throughout this article is based on the following assumption:

Assumption 1.

For any qβ‰₯1q\geq 1, it holds that

supt∈[0,T]𝔼​[|μ​(t,rt)|q]<∞,supt∈[0,T]𝔼​[|ϕ​(t,rt)|q]<∞,\sup_{t\in[0,T]}\mathbb{E}\left[|\mu(t,r_{t})|^{q}\right]<\infty,\quad\sup_{t\in[0,T]}\mathbb{E}\left[|\phi(t,r_{t})|^{q}\right]<\infty,

and the approximate interest rate (r^i​h)i=1,..,T/h(\hat{r}_{ih})_{i=1,..,T/h} satisfies

maxi=1,..,T/h⁑𝔼​[|r^i​hβˆ’ri​h|q]<cq​hq.\max_{i=1,..,T/h}\mathbb{E}\left[|\hat{r}_{ih}-r_{ih}|^{q}\right]<c_{q}h^{q}.
Remark 1.

A typical time-discrete scheme that may satisfy Assumption 1 is the Milstein scheme. Under some standard assumptions on model coefficients (e.g., Lipschitz continuity, linear growth), the convergence rate in the L2L^{2} norm is one.

Lemma 4.1.

Under Assumption 1, we obtain

𝔼​[|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h|q]=O​(hq),βˆ€qβ‰₯2.\mathbb{E}\left[\left|\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{q}\right]=O(h^{q}),\quad\forall q\geq 2.
Proof.

We see that

𝔼​[|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h|q]\displaystyle\mathbb{E}\left[\left|\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{q}\right]
≀cq​𝔼​[|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1ri​h​h|q]+cq​𝔼​[|βˆ‘i=0T/hβˆ’1(r^i​hβˆ’ri​h)​h|q].\displaystyle\leq c_{q}\mathbb{E}\left[\left|\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}r_{ih}h\right|^{q}\right]+c_{q}\mathbb{E}\left[\left|\sum_{i=0}^{T/h-1}(\hat{r}_{ih}-r_{ih})h\right|^{q}\right]. (4)

Let η​(t):=max⁑{l​h:l​h≀t,l=0,1,2,…}\eta(t):=\max\{lh:lh\leq t,l=0,1,2,...\}. For the first term of (4), we have

𝔼​[|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1ri​h​h|q]\displaystyle\mathbb{E}\left[\left|\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}r_{ih}h\right|^{q}\right]
=𝔼​[|∫0Trη​(t)​dtβˆ’βˆ«0Trt​dt|q]\displaystyle=\mathbb{E}\left[\left|\int_{0}^{T}r_{\eta(t)}\mathrm{d}t-\int_{0}^{T}r_{t}\mathrm{d}t\right|^{q}\right]
≀cq​𝔼​[|∫0T(βˆ«Ξ·β€‹(t)tμ​(u,ru)​du)​dt|q]+cq​𝔼​[|∫0T(βˆ«Ξ·β€‹(t)tϕ​(u,ru)​dWu3)​dt|q].\displaystyle\leq c_{q}\mathbb{E}\left[\left|\int_{0}^{T}\left(\int_{\eta(t)}^{t}\mu(u,r_{u})\mathrm{d}u\right)\mathrm{d}t\right|^{q}\right]+c_{q}\mathbb{E}\left[\left|\int_{0}^{T}\left(\int_{\eta(t)}^{t}\phi(u,r_{u})\mathrm{d}W_{u}^{3}\right)\mathrm{d}t\right|^{q}\right]. (5)

An application of the Fubini theorem yields

∫0T(βˆ«Ξ·β€‹(t)tμ​(u,ru)​du)​dt\displaystyle\int_{0}^{T}\left(\int_{\eta(t)}^{t}\mu(u,r_{u})\mathrm{d}u\right)\mathrm{d}t =∫0T(∫uη​(u)+hμ​(u,ru)​dt)​du\displaystyle=\int_{0}^{T}\left(\int_{u}^{\eta(u)+h}\mu(u,r_{u})\mathrm{d}t\right)\mathrm{d}u
=hβ€‹βˆ«0T(1+η​(u)βˆ’uh)​μ​(u,ru)​du.\displaystyle=h\int_{0}^{T}\left(1+\frac{\eta(u)-u}{h}\right)\mu(u,r_{u})\mathrm{d}u.

Hence, it follows that

𝔼​[|∫0T(βˆ«Ξ·β€‹(t)tμ​(u,ru)​du)​dt|q]≀cq​hqβ€‹βˆ«0T𝔼​[|μ​(u,ru)|q]​du=O​(hq).\mathbb{E}\left[\left|\int_{0}^{T}\left(\int_{\eta(t)}^{t}\mu(u,r_{u})\mathrm{d}u\right)\mathrm{d}t\right|^{q}\right]\leq c_{q}h^{q}\int_{0}^{T}\mathbb{E}[|\mu(u,r_{u})|^{q}]\mathrm{d}u=O(h^{q}). (6)

Furthermore, we obtain from the stochastic Fubini theorem (see Theorem 65, Protter [19]), the Burkholder-Davies-Gundy inequality and the Cauchy-Schwarz inequality that

𝔼​[|∫0T(βˆ«Ξ·β€‹(t)tϕ​(u,ru)​dWu3)​dt|q]\displaystyle\mathbb{E}\left[\left|\int_{0}^{T}\left(\int_{\eta(t)}^{t}\phi(u,r_{u})\mathrm{d}W_{u}^{3}\right)\mathrm{d}t\right|^{q}\right]
=hq​𝔼​[|∫0T(1+η​(u)βˆ’uh)​ϕ​(u,ru)​dWu3|q]\displaystyle=h^{q}\mathbb{E}\left[\left|\int_{0}^{T}\left(1+\frac{\eta(u)-u}{h}\right)\phi(u,r_{u})\mathrm{d}W_{u}^{3}\right|^{q}\right]
≀cq​hq​𝔼​[|∫0TΟ•2​(u,ru)​du|q/2]≀cq​hqβ€‹βˆ«0T𝔼​[|ϕ​(u,ru)|q]​du=O​(hq).\displaystyle\leq c_{q}h^{q}\mathbb{E}\left[\left|\int_{0}^{T}\phi^{2}(u,r_{u})\mathrm{d}u\right|^{q/2}\right]\leq c_{q}h^{q}\int_{0}^{T}\mathbb{E}[|\phi(u,r_{u})|^{q}]\mathrm{d}u=O(h^{q}). (7)

Therefore, substituting (6) and (7) into (5), we show that (5) is O​(hq)O(h^{q}). For the second term of (4), by Jensen’s inequality, we have

𝔼​[|βˆ‘i=0T/hβˆ’1(r^i​hβˆ’ri​h)​h|q]≀cqβ€‹βˆ‘i=0T/hβˆ’1𝔼​[|r^i​hβˆ’ri​h|q]​h=O​(hq).\mathbb{E}\left[\left|\sum_{i=0}^{T/h-1}(\hat{r}_{ih}-r_{ih})h\right|^{q}\right]\leq c_{q}\sum_{i=0}^{T/h-1}\mathbb{E}\left[|\hat{r}_{ih}-r_{ih}|^{q}\right]h=O(h^{q}). (8)

Thus, combining (5) and (8) into (4), the proof is complete. ∎

Lemma 4.2.

For the stochastic process (Vt)t∈[0,T](V_{t})_{t\in[0,T]}, we have

𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h|q]=O​(hq),βˆ€qβ‰₯1.\mathbb{E}\left[\left|\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}\right|^{q}\right]=O(h^{q}),\quad\forall q\geq 1.
Proof.

The Cauchy-Schwarz inequality implies that

𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h|q]\displaystyle\mathbb{E}\left[\left|\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}\right|^{q}\right]
=𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h∫0TVt​dt+βˆ‘i=0T/hβˆ’1Vi​h​h|q]\displaystyle=\mathbb{E}\left[\left|\frac{\int_{0}^{T}V_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}V_{ih}h}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}+\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}\right|^{q}\right]
≀𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h∫0TVt​dt|q]\displaystyle\leq\mathbb{E}\left[\left|\frac{\int_{0}^{T}V_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}V_{ih}h}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right|^{q}\right]
≀𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h|2​q]⋅𝔼​[(1∫0TVt​dt)q].\displaystyle\leq\sqrt{\mathbb{E}\left[\left|\int_{0}^{T}V_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}V_{ih}h\right|^{2q}\right]}\cdot\sqrt{\mathbb{E}\left[\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)^{q}\right]}. (9)

From Theorem 4.1(a) in Dufresne [8], we learn that

𝔼​[(1∫0TVt​dt)q]<∞\mathbb{E}\left[\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)^{q}\right]<\infty

for any qβˆˆβ„q\in\mathbb{R}. Since the coefficients in the equation of (Vt)t∈[0,T](V_{t})_{t\in[0,T]} satisfy Assumption 1, we conclude from Lemma 4.1 that 𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h|2​q]=O​(h2​q)\mathbb{E}\left[\left|\int_{0}^{T}V_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}V_{ih}h\right|^{2q}\right]=O(h^{2q}). Therefore, the term of (9) is O​(hq)O(h^{q}) and the proof is complete. ∎

Theorem 4.1.

Under Assumption 1, we have

𝔼​[|XTβˆ’X^Th|q]=O​(hq),βˆ€qβ‰₯2.\mathbb{E}\left[\left|X_{T}-\hat{X}_{T}^{h}\right|^{q}\right]=O(h^{q}),\quad\forall q\geq 2.
Proof.

Straightforward calculation shows that

𝔼​[|XTβˆ’X^Th|q]\displaystyle\mathbb{E}\left[\left|X_{T}-\hat{X}_{T}^{h}\right|^{q}\right]
≀cq​𝔼​[|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h|q]+cq​𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h|q]\displaystyle\leq c_{q}\mathbb{E}\left[\left|\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{q}\right]+c_{q}\mathbb{E}\left[\left|\int_{0}^{T}V_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}V_{ih}h\right|^{q}\right]
+cq​𝔼​[|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h|q].\displaystyle\quad+c_{q}\mathbb{E}\left[\left|\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}\right|^{q}\right].

and an application of Lemmas 4.1 and 4.2 completes the proof. ∎

Then we proceed to the convergence rate of the underlying log-Euler scheme to approximate the price of an option. Different types of options may have different payoff functions, which can be continuous or discontinuous.

4.1 Analysis for continuous payoffs

For continuous payoffs, we impose an assumption:

Assumption 2.

The payoff P:[0,+∞)→ℝP:[0,+\infty)\rightarrow\mathbb{R} is Lipschitz continuous and there exists a constant C>0C>0, such that P​(U)=P​(C)P(U)=P(C) for all U>CU>C.

Under Assumption 2, it holds that

|P​(U1)βˆ’P​(U2)|≀c​|ln⁑U1βˆ’ln⁑U2||P(U_{1})-P(U_{2})|\leq c|\ln U_{1}-\ln U_{2}| (10)

for all U1,U2∈[0,+∞)U_{1},U_{2}\in[0,+\infty), see Theorem 3.1 in Zheng [22]. The payoff that satisfies Assumption 2 is bounded, which is typically suitable for a put-style option. The put-style option becomes worthless when the price of the underlying asset STS_{T} is sufficiently high. For example, the standard European put option has payoff P​(ST):=max⁑{Kβˆ’ST,0}P(S_{T}):=\max\{K-S_{T},0\} with K>0K>0, which satisfies Assumption 2.

Theorem 4.2.

Suppose that Assumptions 1 and 2 are satisfied. Suppose that there exists p>1p>1, such that 𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)<∞\mathbb{E}\left(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t}\right)<\infty and suph𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1r^i​h​h)<∞\sup_{h}\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)<\infty. Then, we have

𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST))2]<∞\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})\right)^{2}\right]<\infty

and

𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST)βˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h​P​(S^Th))2]=O​(h2).\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}P(\hat{S}_{T}^{h})\right)^{2}\right]=O(h^{2}).
Proof.

For the first term, it follows from Jensen’s inequality and the boundedness of P​(ST)P(S_{T}) that

𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST))2]<c​𝔼​(eβˆ’2β€‹βˆ«0Trt​dt)<c​[𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)]1/p<∞.\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})\right)^{2}\right]<c\mathbb{E}\left(e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\right)<c\left[\mathbb{E}\left(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t}\right)\right]^{1/p}<\infty.

Then, we focus on the second term. The Taylor expansion, together with HoΒ¨\ddot{o}lder’s inequality, gives

𝔼​[(eβˆ’βˆ«0Trt​dtβˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h)2]\displaystyle\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)^{2}\right]
=𝔼​[eβˆ’2​Ρ​(∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h)2]\displaystyle=\mathbb{E}\left[e^{-2\varepsilon}\left(\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right)^{2}\right]
≀[𝔼​(max⁑(eβˆ’2​pβ€‹βˆ«0Trt​dt,eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1r^i​h​h))]1/pβ‹…[𝔼​|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h|2​q]1/q\displaystyle\leq\left[\mathbb{E}\left(\max(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t},e^{-2p\sum_{i=0}^{T/h-1}\hat{r}_{ih}h})\right)\right]^{1/p}\cdot\left[\mathbb{E}\left|\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{2q}\right]^{1/q}

where 1p+1q=1\frac{1}{p}+\frac{1}{q}=1, and Ξ΅\varepsilon is between ∫0Trt​dt\int_{0}^{T}r_{t}\mathrm{d}t and βˆ‘i=0T/hβˆ’1r^i​h​h\sum_{i=0}^{T/h-1}\hat{r}_{ih}h. Lemma 4.1 shows that

𝔼​[|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h|2​q]=O​(h2​q).\mathbb{E}\left[\left|\int_{0}^{T}r_{t}\mathrm{d}t-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{2q}\right]=O(h^{2q}).

Note that

𝔼​(max⁑(eβˆ’2​pβ€‹βˆ«0Trt​dt,eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1r^i​h​h))≀𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)+suph𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1r^i​h​h)<c,\mathbb{E}\left(\max(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t},e^{-2p\sum_{i=0}^{T/h-1}\hat{r}_{ih}h})\right)\leq\mathbb{E}\left(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t}\right)+\sup_{h}\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)<c,

where cc is independent of hh. Consequently

𝔼​[(eβˆ’βˆ«0Trt​dtβˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h)2]=O​(h2).\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)^{2}\right]=O(h^{2}).

Therefore, using the boundedness of PP, HoΒ¨\ddot{o}lder’s inequality, inequality (10) and Theorem 4.1, we have

𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST)βˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h​P​(S^Th))2]\displaystyle\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}P(\hat{S}_{T}^{h})\right)^{2}\right]
=𝔼​[(eβˆ’βˆ«0Trt​dt​(P​(ST)βˆ’P​(S^Th))+P​(S^Th)​(eβˆ’βˆ«0Trt​dtβˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h))2]\displaystyle=\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}\left(P(S_{T})-P(\hat{S}_{T}^{h})\right)+P(\hat{S}_{T}^{h})\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)\right)^{2}\right]
≀2​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(P​(ST)βˆ’P​(S^Th))2]+2​𝔼​[P2​(S^Th)​(eβˆ’βˆ«0Trt​dtβˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h)2]\displaystyle\leq 2\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(P(S_{T})-P(\hat{S}_{T}^{h})\right)^{2}\right]+2\mathbb{E}\left[P^{2}(\hat{S}_{T}^{h})\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)^{2}\right]
≀c​[𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)]1/pβ‹…[𝔼​|P​(ST)βˆ’P​(S^Th)|2​q]1/q+c​𝔼​[(eβˆ’βˆ«0Trt​dtβˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h)2]\displaystyle\leq c\left[\mathbb{E}\left(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t}\right)\right]^{1/p}\cdot\left[\mathbb{E}\left|P(S_{T})-P(\hat{S}_{T}^{h})\right|^{2q}\right]^{1/q}+c\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)^{2}\right]
≀c​[𝔼​|ln⁑(ST)βˆ’ln⁑(S^Th)|2​q]1/q+c​𝔼​[(eβˆ’βˆ«0Trt​dtβˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h)2]=O​(h2),\displaystyle\leq c\left[\mathbb{E}\left|\ln(S_{T})-\ln(\hat{S}_{T}^{h})\right|^{2q}\right]^{1/q}+c\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)^{2}\right]=O(h^{2}),

which completes the proof.

∎

Theorem 4.2 is based on Assumption 2 for bounded and Lipschitz payoffs. Fortunately, for unbounded payoffs, such as the European call option with the payoff P​(ST):=max⁑{STβˆ’K,0}P(S_{T}):=\max\{S_{T}-K,0\}, K>0K>0, our numerical experiment in the next section suggests that a similar conclusion to Theorem 4.2 might still hold, but the rigorous convergence analysis is difficult.

4.2 Analysis for digital options

The digital option may be one of the most popular options in finance that has a discontinuous payoff. Next, we extend the analysis to the digital option. The digital call option has the payoff

P​(ST):=1ST>KP(S_{T}):=1_{S_{T}>K}

where K>0K>0. Due to the discontinuity, a direct application of our time-discrete scheme in Section 2 might lead to a slow convergence rate. Therefore, we consider an approach using conditional expectations similar to that in Giles [10] to achieve a better convergence rate. Specifically, conditional on V:=(Vt)t∈[0,T]V:=(V_{t})_{t\in[0,T]} and r:=(rt)t∈[0,T]r:=(r_{t})_{t\in[0,T]}, from equation (1), we have

𝔼​(eβˆ’βˆ«0Trt​dt​1ST>K)\displaystyle\mathbb{E}\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}1_{S_{T}>K}\right)
=𝔼​[𝔼​(eβˆ’βˆ«0Trt​dt​1ST>K|V,r)]\displaystyle=\mathbb{E}\left[\mathbb{E}\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}1_{S_{T}>K}|V,r\right)\right]
=𝔼​[eβˆ’βˆ«0Trt​dt​ℙ​(ST>K|V,r)]\displaystyle=\mathbb{E}\left[e^{-\int_{0}^{T}r_{t}\mathrm{d}t}\mathbb{P}(S_{T}>K|V,r)\right]
=𝔼​[eβˆ’βˆ«0Trt​dt​Φ​(ln⁑S0βˆ’ln⁑K+∫0Trt​dt+(ρ​kΟƒβˆ’12)β€‹βˆ«0TVt​dt+ρσ​(VTβˆ’V0βˆ’k​θ​T)1βˆ’Ο2β€‹βˆ«0TVt​dt)]\displaystyle=\mathbb{E}\left[e^{-\int_{0}^{T}r_{t}\mathrm{d}t}\Phi\left(\frac{\ln S_{0}-\ln K+\int_{0}^{T}r_{t}\mathrm{d}t+\left(\frac{\rho k}{\sigma}-\frac{1}{2}\right)\int_{0}^{T}V_{t}\mathrm{d}t+\frac{\rho}{\sigma}\left(V_{T}-V_{0}-k\theta T\right)}{\sqrt{1-\rho^{2}}\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)\right]
=:𝔼[g(VT,∫0TVtdt,∫0Trtdt)],\displaystyle=:\mathbb{E}\left[g\left(V_{T},\int_{0}^{T}V_{t}\mathrm{d}t,\int_{0}^{T}r_{t}\mathrm{d}t\right)\right], (11)

where Ξ¦\Phi is the cumulative distribution function of the standard normal distribution. Then, we convert the approximation of 𝔼​(eβˆ’βˆ«0Trt​dt​1ST>K)\mathbb{E}\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}1_{S_{T}>K}\right) to the approximation of 𝔼​[g​(VT,∫0TVt​dt,∫0Trt​dt)]\mathbb{E}\left[g\left(V_{T},\int_{0}^{T}V_{t}\mathrm{d}t,\int_{0}^{T}r_{t}\mathrm{d}t\right)\right], where the integrals ∫0TVt​dt\int_{0}^{T}V_{t}\mathrm{d}t and ∫0Trt​dt\int_{0}^{T}r_{t}\mathrm{d}t can be approximated by βˆ‘i=0T/hβˆ’1Vi​h​h\sum_{i=0}^{T/h-1}V_{ih}h and βˆ‘i=0T/hβˆ’1r^i​h​h\sum_{i=0}^{T/h-1}\hat{r}_{ih}h, respectively, based on the Euler scheme.

Lemma 4.3.

Let h=T/nh=T/n, nβˆˆβ„•+n\in\mathbb{N}^{+}. For any qβ‰₯βˆ’2​k​θσ2q\geq-\frac{2k\theta}{\sigma^{2}}, we have

suph𝔼​[(βˆ‘i=0T/hβˆ’1Vi​h​h)q]<∞.\sup_{h}\mathbb{E}\left[\left(\sum_{i=0}^{T/h-1}V_{ih}h\right)^{q}\right]<\infty.
Proof.

It is known that supt∈[0,T]𝔼​(Vtq)<∞\sup_{t\in[0,T]}\mathbb{E}(V_{t}^{q})<\infty for any qβ‰₯βˆ’2​k​θσ2q\geq-\frac{2k\theta}{\sigma^{2}} (see Theorem 3.1, Hurd and Kuznetsov [15] or Dereich, Neuenkirch and Szpruch [7]). If q∈[0,1]q\in[0,1], an application of Jensen’s inequality yields

suph𝔼​[(βˆ‘i=0T/hβˆ’1Vi​h​h)q]≀suph(βˆ‘i=0T/hβˆ’1𝔼​(Vi​h)​h)q≀(T​supt∈[0,T]𝔼​(Vt))q<∞.\sup_{h}\mathbb{E}\left[\left(\sum_{i=0}^{T/h-1}V_{ih}h\right)^{q}\right]\leq\sup_{h}\left(\sum_{i=0}^{T/h-1}\mathbb{E}(V_{ih})h\right)^{q}\leq\left(T\sup_{t\in[0,T]}\mathbb{E}(V_{t})\right)^{q}<\infty.

If q∈[βˆ’2​k​θσ2,0)βˆͺ(1,+∞)q\in[-\frac{2k\theta}{\sigma^{2}},0)\cup(1,+\infty), using Jensen’s inequality again, we have

(1Tβ€‹βˆ‘i=0T/hβˆ’1Vi​h​h)q≀1Tβ€‹βˆ‘i=0T/hβˆ’1Vi​hq​h\left(\frac{1}{T}\sum_{i=0}^{T/h-1}V_{ih}h\right)^{q}\leq\frac{1}{T}\sum_{i=0}^{T/h-1}V_{ih}^{q}h

It follows that

suph𝔼​[(βˆ‘i=0T/hβˆ’1Vi​h​h)q]≀cq​supt∈[0,T]𝔼​(Vtq)<∞.\sup_{h}\mathbb{E}\left[\left(\sum_{i=0}^{T/h-1}V_{ih}h\right)^{q}\right]\leq c_{q}\sup_{t\in[0,T]}\mathbb{E}(V_{t}^{q})<\infty.

The proof is complete. ∎

Remark 2.

Dufresne [8] proved that 𝔼​[(1∫0TVt​dt)βˆ’q]<∞\mathbb{E}\left[\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)^{-q}\right]<\infty for any qβˆˆβ„q\in\mathbb{R}. Lemma 4.3 is consistent with Dufresne’s result, since βˆ‘i=0T/hβˆ’1Vi​h​h\sum_{i=0}^{T/h-1}V_{ih}h converges almost surely to ∫0TVt​dt\int_{0}^{T}V_{t}\mathrm{d}t. However, it seems difficult to prove suph𝔼​[(1βˆ‘i=0T/hβˆ’1Vi​h​h)βˆ’q]<∞\sup_{h}\mathbb{E}\left[\left(\frac{1}{\sum_{i=0}^{T/h-1}V_{ih}h}\right)^{-q}\right]<\infty when q<βˆ’2​k​θσ2q<-\frac{2k\theta}{\sigma^{2}}.

Theorem 4.3.

Let gg be the function defined at (11) associated with the payoff P​(ST)=1ST>KP(S_{T})=1_{S_{T}>K}. Suppose that Assumption 1 is satisfied and that there exists p>1p>1, such that 𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)<∞\mathbb{E}\left(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t}\right)<\infty and suph𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1r^i​h​h)<∞\sup_{h}\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right)<\infty. Let 2​k​θσ2>1\frac{2k\theta}{\sigma^{2}}>1. Then we have

𝔼​[(g​(VT,∫0TVt​dt,∫0Trt​dt)βˆ’g​(VT,βˆ‘i=0T/hβˆ’1Vi​h​h,βˆ‘i=0T/hβˆ’1r^i​h​h))2]=O​(h2).\mathbb{E}\left[\left(g\left(V_{T},\int_{0}^{T}V_{t}\mathrm{d}t,\int_{0}^{T}r_{t}\mathrm{d}t\right)-g\left(V_{T},\sum_{i=0}^{T/h-1}V_{ih}h,\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right)\right)^{2}\right]=O(h^{2}).
Proof.

For notational convenience, we denote by

Ο•:=Φ​(ln⁑S0βˆ’ln⁑K+∫0Trt​dt+(ρ​kΟƒβˆ’12)β€‹βˆ«0TVt​dt+ρσ​(VTβˆ’V0βˆ’k​θ​T)1βˆ’Ο2β€‹βˆ«0TVt​dt)\phi:=\Phi\left(\frac{\ln S_{0}-\ln K+\int_{0}^{T}r_{t}\mathrm{d}t+\left(\frac{\rho k}{\sigma}-\frac{1}{2}\right)\int_{0}^{T}V_{t}\mathrm{d}t+\frac{\rho}{\sigma}\left(V_{T}-V_{0}-k\theta T\right)}{\sqrt{1-\rho^{2}}\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)

and

Ο•^:=Φ​(ln⁑S0βˆ’ln⁑K+βˆ‘i=0T/hβˆ’1r^i​h​h+(ρ​kΟƒβˆ’12)β€‹βˆ‘i=0T/hβˆ’1Vi​h​h+ρσ​(VTβˆ’V0βˆ’k​θ​T)1βˆ’Ο2β€‹βˆ‘i=0T/hβˆ’1Vi​h​h),\hat{\phi}:=\Phi\left(\frac{\ln S_{0}-\ln K+\sum_{i=0}^{T/h-1}\hat{r}_{ih}h+\left(\frac{\rho k}{\sigma}-\frac{1}{2}\right)\sum_{i=0}^{T/h-1}V_{ih}h+\frac{\rho}{\sigma}\left(V_{T}-V_{0}-k\theta T\right)}{\sqrt{1-\rho^{2}}\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}\right),

where Ξ¦\Phi is the cumulative distribution function of the standard normal distribution. Since |Φ​(x)|<1|\Phi(x)|<1 for all xx, we obtain

𝔼​[(g​(VT,∫0TVt​dt,∫0Trt​dt)βˆ’g​(VT,βˆ‘i=0T/hβˆ’1Vi​h​h,βˆ‘i=0T/hβˆ’1r^i​h​h))2]\displaystyle\mathbb{E}\left[\left(g\left(V_{T},\int_{0}^{T}V_{t}\mathrm{d}t,\int_{0}^{T}r_{t}\mathrm{d}t\right)-g\left(V_{T},\sum_{i=0}^{T/h-1}V_{ih}h,\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right)\right)^{2}\right]
≀2​𝔼​[(eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​hβˆ’eβˆ’βˆ«0Trt​dt)2​ϕ^2]+2​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(Ο•^βˆ’Ο•)2]\displaystyle\leq 2\mathbb{E}\left[\left(e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}-e^{-\int_{0}^{T}r_{t}\mathrm{d}t}\right)^{2}\hat{\phi}^{2}\right]+2\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}(\hat{\phi}-\phi)^{2}\right]
≀2​𝔼​[(eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​hβˆ’eβˆ’βˆ«0Trt​dt)2]+2​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(Ο•^βˆ’Ο•)2].\displaystyle\leq 2\mathbb{E}\left[\left(e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}-e^{-\int_{0}^{T}r_{t}\mathrm{d}t}\right)^{2}\right]+2\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}(\hat{\phi}-\phi)^{2}\right]. (12)

For the first term, Theorem 4.2 implies that

𝔼​[(eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​hβˆ’eβˆ’βˆ«0Trt​dt)2]=O​(h2).\mathbb{E}\left[\left(e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}-e^{-\int_{0}^{T}r_{t}\mathrm{d}t}\right)^{2}\right]=O(h^{2}). (13)

For the second term, as the normal cumulative distribution function Ξ¦\Phi is Lipschitz continuous, we have

𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(Ο•^βˆ’Ο•)2]\displaystyle\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}(\hat{\phi}-\phi)^{2}\right]
≀c​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(1βˆ‘i=0T/hβˆ’1Vi​h​hβˆ’1∫0TVt​dt)2]\displaystyle\leq c\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\frac{1}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}-\frac{1}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)^{2}\right]
+c​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(VTβˆ‘i=0T/hβˆ’1Vi​h​hβˆ’VT∫0TVt​dt)2]\displaystyle\quad+c\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\frac{V_{T}}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}-\frac{V_{T}}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)^{2}\right]
+c​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(βˆ‘i=0T/hβˆ’1Vi​h​hβˆ’βˆ«0TVt​dt)2]\displaystyle\quad+c\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}-\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}\right)^{2}\right]
+c​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(βˆ‘i=0T/hβˆ’1r^i​h​hβˆ‘i=0T/hβˆ’1Vi​h​hβˆ’βˆ«0Trt​dt∫0TVt​dt)2].\displaystyle\quad+c\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\frac{\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}-\frac{\int_{0}^{T}r_{t}\mathrm{d}t}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)^{2}\right]. (14)

Let’s focus on the last term of (14). Since

|βˆ‘i=0T/hβˆ’1r^i​h​hβˆ‘i=0T/hβˆ’1Vi​h​hβˆ’βˆ«0Trt​dt∫0TVt​dt|\displaystyle\left|\frac{\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}-\frac{\int_{0}^{T}r_{t}\mathrm{d}t}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right|
≀|βˆ‘i=0T/hβˆ’1r^i​h​h|​|1βˆ‘i=0T/hβˆ’1Vi​h​hβˆ’1∫0TVt​dt|+1∫0TVt​dt​|βˆ‘i=0T/hβˆ’1r^i​h​hβˆ’βˆ«0Trt​dt|\displaystyle\leq\left|\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|\left|\frac{1}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}-\frac{1}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right|+\frac{1}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\left|\sum_{i=0}^{T/h-1}\hat{r}_{ih}h-\int_{0}^{T}r_{t}\mathrm{d}t\right|
=|βˆ‘i=0T/hβˆ’1r^i​h​h|​|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​hβˆ‘i=0T/hβˆ’1Vi​h​hβ€‹βˆ«0TVt​dt|+1∫0TVt​dt​|βˆ‘i=0T/hβˆ’1r^i​h​hβˆ’βˆ«0Trt​dt|,\displaystyle=\left|\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|\left|\frac{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right|+\frac{1}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\left|\sum_{i=0}^{T/h-1}\hat{r}_{ih}h-\int_{0}^{T}r_{t}\mathrm{d}t\right|,

using the fact that the stochastic processes rr and r^\hat{r} are independent of VV, we obtain

𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(βˆ‘i=0T/hβˆ’1r^i​h​hβˆ‘i=0T/hβˆ’1Vi​h​hβˆ’βˆ«0Trt​dt∫0TVt​dt)2]\displaystyle\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\frac{\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}-\frac{\int_{0}^{T}r_{t}\mathrm{d}t}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)^{2}\right]
≀c​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(βˆ‘i=0T/hβˆ’1r^i​h​h)2]​𝔼​[(∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​hβˆ‘i=0T/hβˆ’1Vi​h​hβ€‹βˆ«0TVt​dt)2]\displaystyle\leq c\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right)^{2}\right]\mathbb{E}\left[\left(\frac{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)^{2}\right]
+c​𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(βˆ‘i=0T/hβˆ’1r^i​h​hβˆ’βˆ«0Trt​dt)2]​𝔼​[(1∫0TVt​dt)].\displaystyle\quad+c\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\sum_{i=0}^{T/h-1}\hat{r}_{ih}h-\int_{0}^{T}r_{t}\mathrm{d}t\right)^{2}\right]\mathbb{E}\left[\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)\right].

Recall that 𝔼​[(1∫0TVt​dt)q]<∞\mathbb{E}\left[\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)^{q}\right]<\infty for any qβˆˆβ„q\in\mathbb{R}. Furthermore, by Jensen’s inequality, it holds that

𝔼​[|βˆ‘i=0T/hβˆ’1r^i​h​h|q]≀cq​𝔼​(βˆ‘i=0T/hβˆ’1|r^i​h|q​h)≀cq​maxi=1,..,T/h⁑𝔼​(|r^i​h|q)<cq\mathbb{E}\left[\left|\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{q}\right]\leq c_{q}\mathbb{E}\left(\sum_{i=0}^{T/h-1}|\hat{r}_{ih}|^{q}h\right)\leq c_{q}\max_{i=1,..,T/h}\mathbb{E}\left(|\hat{r}_{ih}|^{q}\right)<c_{q}

for any qβ‰₯1q\geq 1, where cqc_{q} is independent of hh. Therefore, applying HoΒ¨\ddot{o}lder’s inequality, together with Lemma 4.3, we conclude that

𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(βˆ‘i=0T/hβˆ’1r^i​h​hβˆ‘i=0T/hβˆ’1Vi​h​hβˆ’βˆ«0Trt​dt∫0TVt​dt)2]\displaystyle\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}\left(\frac{\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}{\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}-\frac{\int_{0}^{T}r_{t}\mathrm{d}t}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right)^{2}\right]
≀c​[𝔼​|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h∫0TVt​dt|2​q1]1/q1​[𝔼​(1βˆ‘i=0T/hβˆ’1Vi​h​h)p1]1/p1\displaystyle\leq c\left[\mathbb{E}\left|\frac{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}}{\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}}\right|^{2q_{1}}\right]^{1/q_{1}}\left[\mathbb{E}\left(\frac{1}{\sum_{i=0}^{T/h-1}V_{ih}h}\right)^{p_{1}}\right]^{1/p_{1}}
Γ—[𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)]1/p​[𝔼​|βˆ‘i=0T/hβˆ’1r^i​h​h|2​q]1/q\displaystyle\quad\times\left[\mathbb{E}(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t})\right]^{1/p}\left[\mathbb{E}\left|\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{2q}\right]^{1/q}
+c​[𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)]1/p​[𝔼​|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h|2​q]1/q​𝔼​(1∫0TVt​dt)\displaystyle\quad+c\left[\mathbb{E}(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t})\right]^{1/p}\left[\mathbb{E}\left|\sqrt{\int_{0}^{T}r_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right|^{2q}\right]^{1/q}\mathbb{E}\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)
≀c​[𝔼​|∫0TVt​dtβˆ’βˆ‘i=0T/hβˆ’1Vi​h​h|4​q1​𝔼​(1∫0TVt​dt)2​q1]1/(2​q1)​[𝔼​(1βˆ‘i=0T/hβˆ’1Vi​h​h)p1]1/p1\displaystyle\leq c\left[\mathbb{E}\left|\sqrt{\int_{0}^{T}V_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}V_{ih}h}\right|^{4q_{1}}\mathbb{E}\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)^{2q_{1}}\right]^{1/(2q_{1})}\left[\mathbb{E}\left(\frac{1}{\sum_{i=0}^{T/h-1}V_{ih}h}\right)^{p_{1}}\right]^{1/p_{1}}
Γ—[𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)]1/p​[𝔼​|βˆ‘i=0T/hβˆ’1r^i​h​h|2​q]1/q\displaystyle\quad\times\left[\mathbb{E}(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t})\right]^{1/p}\left[\mathbb{E}\left|\sum_{i=0}^{T/h-1}\hat{r}_{ih}h\right|^{2q}\right]^{1/q}
+c​[𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)]1/p​[𝔼​|∫0Trt​dtβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h|2​q]1/q​𝔼​(1∫0TVt​dt)=O​(h2),\displaystyle\quad+c\left[\mathbb{E}(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t})\right]^{1/p}\left[\mathbb{E}\left|\sqrt{\int_{0}^{T}r_{t}\mathrm{d}t}-\sqrt{\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}\right|^{2q}\right]^{1/q}\mathbb{E}\left(\frac{1}{\int_{0}^{T}V_{t}\mathrm{d}t}\right)=O(h^{2}),

where p,q,p1,q1p,q,p_{1},q_{1} satisfy 1p+1q=1\frac{1}{p}+\frac{1}{q}=1 and 1p1+1q1=1\frac{1}{p_{1}}+\frac{1}{q_{1}}=1. Note that when 2​k​θσ2>1\frac{2k\theta}{\sigma^{2}}>1, suph𝔼​[(1βˆ‘i=0T/hβˆ’1Vi​h​h)p1]<∞\sup_{h}\mathbb{E}\left[\left(\frac{1}{\sum_{i=0}^{T/h-1}V_{ih}h}\right)^{p_{1}}\right]<\infty (i.e., p1<2​k​θσ2p_{1}<\frac{2k\theta}{\sigma^{2}}, see Lemma 4.3) can be achieved by choosing p1p_{1} sufficiently close to 11. The other terms of (14) can be shown to be O​(h2)O(h^{2}) analogously, hence it holds that 𝔼​[eβˆ’2β€‹βˆ«0Trt​dt​(Ο•^βˆ’Ο•)2]=O​(h2)\mathbb{E}\left[e^{-2\int_{0}^{T}r_{t}\mathrm{d}t}(\hat{\phi}-\phi)^{2}\right]=O(h^{2}). Combining this estimate and (13) into (12), we complete the proof. ∎

The analysis above is for the digital call option. For the digital put option with payoff P​(ST):=1K>STP(S_{T}):=1_{K>S_{T}}, analogous convergence results can be easily derived following analogous techniques and proofs.

5 Applications

In this section, we apply our results from Section 4 to several well-known interest rate models in finance, including the CIR model, the Hull-White model and the Black-Karasinski model. The option payoff PP we consider satisfies either Assumption 2 or is the payoff of a digital option.

5.1 Heston-CIR model

The CIR model, which was introduced by Cox, Ingersoll and Ross [5], is represented as

d​rt=α​(Ξ²βˆ’rt)​d​t+γ​rt​d​Wt3,\mathrm{d}r_{t}=\alpha(\beta-r_{t})\mathrm{d}t+\gamma\sqrt{r_{t}}\mathrm{d}W_{t}^{3},

where α,β,γ,r0>0\alpha,\beta,\gamma,r_{0}>0. It is known that rtr_{t} follows a scaled noncentral chi-squared distribution given ru,u∈[0,t)r_{u},u\in[0,t), i.e.,

rt​=d​γ2​(1βˆ’eβˆ’Ξ±β€‹(tβˆ’u))4​α​χd2​(4​α​eβˆ’k​(tβˆ’u)Οƒ2​(1βˆ’eβˆ’Ξ±β€‹(tβˆ’u))​ru),r_{t}\overset{\mathrm{d}}{=}\frac{\gamma^{2}(1-\mathrm{e}^{-\alpha(t-u)})}{4\alpha}\chi_{d}^{2}\left(\frac{4\alpha\mathrm{e}^{-k(t-u)}}{\sigma^{2}(1-\mathrm{e}^{-\alpha(t-u)})}r_{u}\right),

where Ο‡d2​(Ξ»)\chi_{d}^{2}(\lambda) denotes a noncentral chi-squared random variable with degrees of freedom d=4​α​βγ2d=\frac{4\alpha\beta}{\gamma^{2}} and noncentrality parameter Ξ»\lambda (see Glasserman [11]). For the exact simulation of rtr_{t}, i.e., r^t=rt\hat{r}_{t}=r_{t}, t∈[0,T]t\in[0,T], we have supt∈[0,T]𝔼​(rtq)<∞\sup_{t\in[0,T]}\mathbb{E}(r_{t}^{q})<\infty for any q>βˆ’2​α​βγ2q>-\frac{2\alpha\beta}{\gamma^{2}}. Hence, it is easy to verify that Assumption 1 is satisfied. As ℙ​(rtβ‰₯0,βˆ€t∈[0,T])=1\mathbb{P}(r_{t}\geq 0,\forall t\in[0,T])=1, the moment condition in Theorem 4.2 is satisfied and it follows from Theorem 4.2 that

𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST)βˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h​P​(S^Th))2]=O​(h2).\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}P(\hat{S}_{T}^{h})\right)^{2}\right]=O(h^{2}). (15)

Note that Theorem 4.3 has the same assumptions as Theorem 4.2. Hence, for digital options, equation (15) also holds.

Since the exact simulation of the CIR process can be time-consuming, there are several time-discrete schemes for the CIR process (see Alfonsi [1] for discussions). Neuenkirch and Szpruch [18] showed that the BEM scheme and the drift-implicit Milstein scheme preserve the nonnegativity of the CIR process, i.e., ℙ​(r^tβ‰₯0,βˆ€t∈[0,T])=1\mathbb{P}(\hat{r}_{t}\geq 0,\forall t\in[0,T])=1 and both are strongly convergent with order one when 2​α​βγ2>3\frac{2\alpha\beta}{\gamma^{2}}>3. Specifically, the BEM scheme can be written as

x^i​h=x^(iβˆ’1)​h+Ξ±2​((Ξ²βˆ’Ξ³24​α)​x^i​hβˆ’1βˆ’x^i​h)​h+Ξ³2​(Wi​hβˆ’W(iβˆ’1)​h),\hat{x}_{ih}=\hat{x}_{(i-1)h}+\frac{\alpha}{2}\left(\left(\beta-\frac{\gamma^{2}}{4\alpha}\right)\hat{x}_{ih}^{-1}-\hat{x}_{ih}\right)h+\frac{\gamma}{2}(W_{ih}-W_{(i-1)h}),

where r^i​h=x^i​h2\hat{r}_{ih}=\hat{x}_{ih}^{2}. Proposition 3.1 in Neuenkirch and Szpruch [18] demonstrated that

𝔼[maxi=1,..,T/h(r^i​hβˆ’ri​h)p]<cnhp,\mathbb{E}\left[\max_{i=1,..,T/h}(\hat{r}_{ih}-r_{ih})^{p}\right]<c_{n}h^{p},

if 2≀p<43​α​βγ22\leq p<\frac{4}{3}\frac{\alpha\beta}{\gamma^{2}}. The drift-implicit Milstein scheme is

r^i​h=r^(iβˆ’1)​h+α​(Ξ²βˆ’r^i​h)​h+γ​r^(iβˆ’1)​h​(Wi​hβˆ’W(iβˆ’1)​h)+Ξ³24​((Wi​hβˆ’W(iβˆ’1)​h)2βˆ’h).\hat{r}_{ih}=\hat{r}_{(i-1)h}+\alpha(\beta-\hat{r}_{ih})h+\gamma\sqrt{\hat{r}_{(i-1)h}}(W_{ih}-W_{(i-1)h})+\frac{\gamma^{2}}{4}((W_{ih}-W_{(i-1)h})^{2}-h).

Lemma 4.1 in Neuenkirch and Szpruch [18] guaranteed that

maxi=1,..,T/h⁑𝔼​|r^i​hβˆ’ri​h|<c​h,\max_{i=1,..,T/h}\mathbb{E}\left|\hat{r}_{ih}-r_{ih}\right|<ch,

if α​βγ2>32\frac{\alpha\beta}{\gamma^{2}}>\frac{3}{2}. These results indicate that Assumption 1 might be satisfied; thus, (15) might hold for both the BEM scheme and the drift-implicit Milstein scheme applied to the CIR process.

5.2 Heston-Hull-White model

The Hull-White model (Hull and White [14]) is of the form

d​rt=α​(β​(t)βˆ’rt)​d​t+γ​d​Wt3,\mathrm{d}r_{t}=\alpha(\beta(t)-r_{t})\mathrm{d}t+\gamma\mathrm{d}W_{t}^{3}, (16)

where Ξ±,Ξ³>0\alpha,\gamma>0, r0βˆˆβ„r_{0}\in\mathbb{R} and Ξ²:[0,T]→ℝ+\beta:[0,T]\rightarrow\mathbb{R}^{+} is continuous. Given ru,u∈[0,t)r_{u},u\in[0,t), the interest rate rtr_{t} is normally distributed with mean

eβˆ’Ξ±β€‹(tβˆ’u)​ru+Ξ±β€‹βˆ«uteβˆ’Ξ±β€‹(tβˆ’s)​β​(s)​dse^{-\alpha(t-u)}r_{u}+\alpha\int_{u}^{t}e^{-\alpha(t-s)}\beta(s)\mathrm{d}s

and variance

Ξ³22​α​(1βˆ’eβˆ’2​α​(tβˆ’u))\frac{\gamma^{2}}{2\alpha}(1-e^{-2\alpha(t-u)})

(see Glasserman [11], p109). In practice, it is often the case that β​(t)\beta(t) has a simple structure so that it is convenient to simulate rtr_{t} exactly. Let r^t=rt\hat{r}_{t}=r_{t}, t∈[0,T]t\in[0,T]. Since it holds that supt∈[0,T]𝔼​[|rt|q]<∞\sup_{t\in[0,T]}\mathbb{E}[|r_{t}|^{q}]<\infty for any qβ‰₯0q\geq 0, Assumption 1 is satisfied. Furthermore, we have 𝔼​(eβˆ’2​pβ€‹βˆ«0Trt​dt)<∞\mathbb{E}\left(e^{-2p\int_{0}^{T}r_{t}\mathrm{d}t}\right)<\infty for any pβˆˆβ„p\in\mathbb{R} (see Glasserman [11], p111). The lemma below shows the boundedness of suph𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1ri​h​h)\sup_{h}\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-1}r_{ih}h}\right).

Lemma 5.1.

Let (rt)t∈[0,T](r_{t})_{t\in[0,T]} satisfy the Hull-White model (16). Let h=T/nh=T/n, nβˆˆβ„•+n\in\mathbb{N}^{+}. For any pβˆˆβ„p\in\mathbb{R}, we have

suph𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1ri​h​h)<∞.\sup_{h}\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-1}r_{ih}h}\right)<\infty.
Proof.

It is easy to show that there exists a constant cc, which is independent of ii and hh, such that

max⁑(|2​pβ€‹Ξ±β€‹βˆ«(iβˆ’1)​hi​heβˆ’Ξ±β€‹(i​hβˆ’s)​β​(s)​ds|,|p2​γ2α​(1βˆ’eβˆ’2​α​h)|)<c​h.\max\left(\left|2p\alpha\int_{(i-1)h}^{ih}e^{-\alpha(ih-s)}\beta(s)\mathrm{d}s\right|,\left|\frac{p^{2}\gamma^{2}}{\alpha}(1-e^{-2\alpha h})\right|\right)<ch.

Moreover, it is known that if XX is normally distributed with mean ΞΌ\mu and variance Οƒ2\sigma^{2}, then 𝔼​(eX)=eΞΌ+12​σ2\mathbb{E}(e^{X})=e^{\mu+\frac{1}{2}\sigma^{2}}. Let

h1\displaystyle h_{1} :=h≀h\displaystyle:=h\leq h
h2\displaystyle h_{2} :=h+eβˆ’Ξ±β€‹h​h1≀2​h\displaystyle:=h+e^{-\alpha h}h_{1}\leq 2h
…\displaystyle...
hi\displaystyle h_{i} :=h+eβˆ’Ξ±β€‹h​hiβˆ’1≀i​h\displaystyle:=h+e^{-\alpha h}h_{i-1}\leq ih
…\displaystyle...
hn\displaystyle h_{n} :=h+eβˆ’Ξ±β€‹h​hnβˆ’1≀T\displaystyle:=h+e^{-\alpha h}h_{n-1}\leq T

where n=T/hn=T/h and Ξ±>0\alpha>0. Then, we have

ln⁑𝔼​(eβˆ’2​p​hi​r(nβˆ’i)​h|r(nβˆ’iβˆ’1)​h)\displaystyle\ln\mathbb{E}(e^{-2ph_{i}r_{(n-i)h}}|r_{(n-i-1)h})
β‰€βˆ’2​p​hi​eβˆ’Ξ±β€‹hβ‹…r(nβˆ’iβˆ’1)​h+c​hβ‹…hi+c​hβ‹…hi2\displaystyle\leq-2ph_{i}e^{-\alpha h}\cdot r_{(n-i-1)h}+ch\cdot h_{i}+ch\cdot h_{i}^{2}
β‰€βˆ’2​p​hi​eβˆ’Ξ±β€‹hβ‹…r(nβˆ’iβˆ’1)​h+c​i​h2+c​i2​h3\displaystyle\leq-2ph_{i}e^{-\alpha h}\cdot r_{(n-i-1)h}+cih^{2}+ci^{2}h^{3}

for any i=1,2,…,nβˆ’1i=1,2,...,n-1. Therefore, based on conditional expectations, we can calculate recursively that

𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1ri​h​h)\displaystyle\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-1}r_{ih}h}\right)
≀𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’2ri​h​hβ‹…eβˆ’2​p​h1​eβˆ’Ξ±β€‹h​r(T/hβˆ’2)​h)​ec​h2+c​h3\displaystyle\leq\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-2}r_{ih}h}\cdot e^{-2ph_{1}e^{-\alpha h}r_{(T/h-2)h}}\right)e^{ch^{2}+ch^{3}}
≀𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’3ri​h​hβ‹…eβˆ’2​p​h2​eβˆ’Ξ±β€‹h​r(T/hβˆ’3)​h)​ecβ€‹βˆ‘i=12i​h2+cβ€‹βˆ‘i=12i2​h3\displaystyle\leq\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-3}r_{ih}h}\cdot e^{-2ph_{2}e^{-\alpha h}r_{(T/h-3)h}}\right)e^{c\sum_{i=1}^{2}ih^{2}+c\sum_{i=1}^{2}i^{2}h^{3}}
≀…\displaystyle\leq...
≀(eβˆ’2​p​r0​hβ‹…eβˆ’2​p​hT/hβˆ’1​eβˆ’Ξ±β€‹h​r0)​ecβ€‹βˆ‘i=1T/hβˆ’1i​h2+cβ€‹βˆ‘i=1T/hβˆ’1i2​h3\displaystyle\leq\left(e^{-2pr_{0}h}\cdot e^{-2ph_{T/h-1}e^{-\alpha h}r_{0}}\right)e^{c\sum_{i=1}^{T/h-1}ih^{2}+c\sum_{i=1}^{T/h-1}i^{2}h^{3}}
=eβˆ’2​p​r0​hT/h​ecβ€‹βˆ‘i=1T/hβˆ’1i​h2+cβ€‹βˆ‘i=1T/hβˆ’1i2​h3≀cp,\displaystyle=e^{-2pr_{0}h_{T/h}}e^{c\sum_{i=1}^{T/h-1}ih^{2}+c\sum_{i=1}^{T/h-1}i^{2}h^{3}}\leq c_{p},

where cpc_{p} is independent of hh. This means that suph𝔼​(eβˆ’2​pβ€‹βˆ‘i=0T/hβˆ’1ri​h​h)\sup_{h}\mathbb{E}\left(e^{-2p\sum_{i=0}^{T/h-1}r_{ih}h}\right) is also bounded and the proof is complete. ∎

Then, all assumptions in Theorems 4.2 or 4.3 are satisfied, and we conclude that

𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST)βˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h​P​(S^Th))2]=O​(h2).\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}P(\hat{S}_{T}^{h})\right)^{2}\right]=O(h^{2}).

5.3 Heston-Black-Karasinski model

The Black-Karasinski model (Black and Karasinski [3]) can be written as

d​ln⁑rt=(β​(t)βˆ’Ξ±β€‹ln⁑rt)​d​t+γ​d​Wt3,\mathrm{d}\ln r_{t}=(\beta(t)-\alpha\ln r_{t})\mathrm{d}t+\gamma\mathrm{d}W_{t}^{3},

where Ξ±,Ξ³,r0>0\alpha,\gamma,r_{0}>0 and Ξ²:[0,T]→ℝ+\beta:[0,T]\rightarrow\mathbb{R}^{+} is continuous. It follows from Ito^\hat{o}’s formula that

d​rt=rt​(β​(t)+Ξ³22βˆ’Ξ±β€‹ln⁑rt)​d​t+γ​rt​d​Wt3.\mathrm{d}r_{t}=r_{t}\left(\beta(t)+\frac{\gamma^{2}}{2}-\alpha\ln r_{t}\right)\mathrm{d}t+\gamma r_{t}\mathrm{d}W_{t}^{3}.

Given ru,u∈[0,t)r_{u},u\in[0,t), the random variable rtr_{t} has a lognormal distribution (see Brigo and Mercurio [4]); hence, rtr_{t} is usually simulated exactly. Specifically, given ru,u∈[0,t)r_{u},u\in[0,t), the logarithm of the interest rate ln⁑rt\ln r_{t} is normally distributed with mean

eβˆ’Ξ±β€‹(tβˆ’u)​ln⁑ru+∫uteβˆ’Ξ±β€‹(tβˆ’s)​β​(s)​dse^{-\alpha(t-u)}\ln r_{u}+\int_{u}^{t}e^{-\alpha(t-s)}\beta(s)\mathrm{d}s

and variance

Ξ³22​α​(1βˆ’eβˆ’2​α​(tβˆ’u)).\frac{\gamma^{2}}{2\alpha}(1-e^{-2\alpha(t-u)}).

Let r^t=rt\hat{r}_{t}=r_{t}, t∈[0,T]t\in[0,T]. Since each moment of a lognormal random variable is finite, we have supt∈[0,T]𝔼​(rtq)<∞\sup_{t\in[0,T]}\mathbb{E}(r_{t}^{q})<\infty for any qβˆˆβ„q\in\mathbb{R}. It holds from the continuity of Ξ²\beta and the Cauchy-Schwarz inequality that

supt∈[0,T]𝔼​[|rt​(β​(t)+Ξ³22βˆ’Ξ±β€‹ln⁑rt)|q]\displaystyle\sup_{t\in[0,T]}\mathbb{E}\left[\left|r_{t}\left(\beta(t)+\frac{\gamma^{2}}{2}-\alpha\ln r_{t}\right)\right|^{q}\right]
≀cq​supt∈[0,T]𝔼​(rtq)+cq​supt∈[0,T]𝔼​[|rt​ln⁑rt|q]\displaystyle\leq c_{q}\sup_{t\in[0,T]}\mathbb{E}(r_{t}^{q})+c_{q}\sup_{t\in[0,T]}\mathbb{E}\left[\left|r_{t}\ln r_{t}\right|^{q}\right]
≀cq​supt∈[0,T]𝔼​(rtq)+cq​supt∈[0,T]𝔼​(rt2​q)β‹…supt∈[0,T]𝔼​[|ln⁑rt|2​q]<∞,\displaystyle\leq c_{q}\sup_{t\in[0,T]}\mathbb{E}(r_{t}^{q})+c_{q}\sqrt{\sup_{t\in[0,T]}\mathbb{E}(r_{t}^{2q})\cdot\sup_{t\in[0,T]}\mathbb{E}\left[\left|\ln r_{t}\right|^{2q}\right]}<\infty,

for any qβ‰₯1q\geq 1. Thus, Assumption 1 is satisfied. As rtr_{t} is nonnegative, the moment condition is automatically satisfied, and we obtain from Theorems 4.2 or 4.3 that

𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST)βˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h​P​(S^Th))2]=O​(h2).\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}P(\hat{S}_{T}^{h})\right)^{2}\right]=O(h^{2}).

6 Numerical results

In this section, we conduct numerical experiments to verify the convergence rate derived in Section 4 and then evaluate the efficiency of the log-Euler scheme we develop combined with Rhee and Glynn’s unbiased estimators.

We consider the Heston model with three different interest rate models: the CIR model, the Hull-White model and the Black-Karasinski model. For the CIR model, we focus on two methods to simulate the path: the exact simulation method and the BEM scheme from Neuenkirch and Szpruch [18]. For the Hull-White model and the Black-Karasinski model, we simulate the paths exactly. We evaluate the convergence rate of the error

E​r​r​(h):=𝔼​[(eβˆ’βˆ«0Trt​dt​P​(ST)βˆ’eβˆ’βˆ‘i=0T/hβˆ’1r^i​h​h​P​(S^Th))2],Err(h):=\mathbb{E}\left[\left(e^{-\int_{0}^{T}r_{t}\mathrm{d}t}P(S_{T})-e^{-\sum_{i=0}^{T/h-1}\hat{r}_{ih}h}P(\hat{S}_{T}^{h})\right)^{2}\right],

with respect to the step size hh, where P​(ST)P(S_{T}) is the payoff of an option. Three types of options are considered:

  • β€’

    European put option P​(ST):=max⁑(Kβˆ’ST,0)P(S_{T}):=\max(K-S_{T},0);

  • β€’

    European call option P​(ST):=max⁑(STβˆ’K,0)P(S_{T}):=\max(S_{T}-K,0);

  • β€’

    Digital call option P​(ST):=1ST>KP(S_{T}):=1_{S_{T}>K}.

The model parameters are available in Table 1 and we set T=1T=1 and S0=K=1S_{0}=K=1 for all cases. These options are interesting for the reason below: a European put option has a bounded and continuous payoff, a European call options has an unbounded payoff and a digital option has a discontinuous payoff. All experiments are performed in Matlab.

Note that the payoff of a European put option satisfies Assumption 2, and according to Theorem 4.2, the theoretical convergence rate should be 22. For the digital call option, Theorem 4.3 guarantees that the rate is also 22 if the approach of conditional expectations is used. The European call option is out of the scope of our analysis.

kk ΞΈ\theta Οƒ\sigma ρ\rho Ξ±\alpha Ξ²\beta Ξ³\gamma r0r_{0} V0V_{0} S0S_{0}
CIR-exact 2.82.8 0.050.05 0.250.25 0.50.5 1.21.2 0.060.06 0.250.25 0.050.05 0.040.04 11
CIR-BEM 33 0.040.04 0.250.25 0.50.5 3.53.5 0.060.06 0.250.25 0.050.05 0.040.04 11
HW 2.82.8 0.050.05 0.250.25 0.50.5 1.21.2 0.060.06 0.50.5 0.050.05 0.040.04 11
BK 2.82.8 0.050.05 0.250.25 0.50.5 1.21.2 0.060.06 0.250.25 0.050.05 0.040.04 11
Table 1: Parameters of the Heston model with stochastic interest rates, where CIR-exact, CIR-BEM, HW and BK represent the CIR model with exact simulation, the CIR model simulated through the BEM method, the Hull-White model and the Black-Karasinski model, respectively.

Figure 1 plots log2⁑(E​r​r​(h))\log_{2}(Err(h)) against βˆ’log2⁑(h)-\log_{2}(h), with h=2βˆ’nh=2^{-n}, n=0,1,2,…,7n=0,1,2,...,7. Here, the exact values of STS_{T} and ∫0Trt​dt\int_{0}^{T}r_{t}\mathrm{d}t are approximated by using the Euler scheme in Section 2 based on a very small step size 2βˆ’102^{-10}. For the BEM method, the exact path of (rt)t∈[0,T](r_{t})_{t\in[0,T]} needs an additional approximation also with step size 2βˆ’102^{-10}, which shares the same Brownian motion path with the corresponding (r^i​h)i=1,2,…,T/h(\hat{r}_{ih})_{i=1,2,...,T/h}. To estimate E​r​r​(h)Err(h), the number of Monte Carlo samples for each quantity in each model is at least 0.50.5 million, so that the standard deviation of the estimator of E​r​r​(h)Err(h) is typically less than 1%1\% of the estimated value of E​r​r​(h)Err(h). As illustrated in Figure 1, the convergence rate in all cases is 22, for all models and all option payoffs that we consider, which is consistent with the theoretical convergence rate. Furthermore, although the European call option does not fall into the scope of our analysis, it still has a desired convergence rate. This implies that our convergence result might be extended for more types of options.

Refer to caption
Refer to caption
Refer to caption
Figure 1: Convergence rate for the Heston model with stochastic interest rates. The model parameters are from Table 1.

Next, we incorporate the log-Euler scheme into Rhee and Glynn’s unbiased estimators. As discussed in Section 3, the implementation of the unbiased estimator ZZ requires setting a distribution of 𝒩\mathcal{N}. In this experiment, we simply take ℙ​(𝒩β‰₯n)=2βˆ’3​n/2\mathbb{P}(\mathcal{N}\geq n)=2^{-3n/2}, nβˆˆβ„•n\in\mathbb{N}, such that (2) and (3) are finite. Hence, ZZ is unbiased with a finite variance and finite computational time. Table 2 reports the root mean square error (RMSE) and the computational time (in seconds) of ZZ based on 11 million samples. For some applications, either the variance or the computational time of ZZ can be infinite; see Zheng, Blanchet and Glynn [25]. We see from Table 2 that all these quantities are finite, which again coincides with the theory. In particular, the computational time for all three options is finite, but due to space limitation, we only illustrate that for the European put option. Therefore, the log-Euler scheme we develop is well-suited to the framework of Rhee and Glynn’s unbiased estimators. Moreover, compared with the result from Figure 1, we observe that a method for an interest rate model with a large RMSE of ZZ tends to have a large E​r​r​(h)Err(h). For example, for the European call option, CIR-BEM has the largest RMSE and E​r​r​(h)Err(h), then follows by BK and HW, and CIR- exact has the smallest errors. On the other hand, the digital option generally has a larger RMSE compared with the European call or put options. This may be because that the payoff function of the digital option is discontinuous at ST=KS_{T}=K, and |P​(S^Th)βˆ’P​(ST)|=1|P(\hat{S}^{h}_{T})-P(S_{T})|=1 can occur when STβ‰ˆKS_{T}\approx K.

RMSE-put RMSE-call RMSE-digital Comput time
CIR-exact 1.21Γ—10βˆ’41.21\times 10^{-4} 3.39Γ—10βˆ’43.39\times 10^{-4} 4.22Γ—10βˆ’44.22\times 10^{-4} 8.788.78
CIR-BEM 1.96Γ—10βˆ’41.96\times 10^{-4} 0.00140.0014 4.45Γ—10βˆ’44.45\times 10^{-4} 8.378.37
HW 4.16Γ—10βˆ’44.16\times 10^{-4} 4.57Γ—10βˆ’44.57\times 10^{-4} 6.32Γ—10βˆ’46.32\times 10^{-4} 5.725.72
BK 1.78Γ—10βˆ’41.78\times 10^{-4} 5.31Γ—10βˆ’45.31\times 10^{-4} 8.32Γ—10βˆ’48.32\times 10^{-4} 6.176.17
Table 2: The RMSE and computational time (in seconds, for the European put option) of ZZ based on 10610^{6} samples. The model parameters are from Table 1.

7 Conclusion

In this article, we develop a semi-exact log-Euler scheme for the Heston model with stochastic interest rates and we analyse the relevant convergence rate in the L2L^{2} norm. The SDEs of the Heston model with stochastic interest rates can be divided into two components: the Heston component and the interest rate component. Under mild assumptions on the interest rate component, we show that the convergence rate in the L2L^{2} norm is one, which enables us to easily incorporate the log-Euler scheme into Rhee and Glynn’s unbiased estimators. In our analysis, we consider two types of payoffs of options: the bounded and Lipschitz payoff and the payoff of a digital option. Furthermore, we demonstrate that the log-Euler scheme and the convergence analysis apply to a large class of interest rate models, including the well-known CIR model, Hull-White model and Black-Karasinski model.

There are two directions of extensions that might be interesting: the log-Euler scheme we consider is based on the assumption that the driven Brownian motion W3W^{3} for the interest rate model is independent of the driven Brownian motions W1W^{1} and W2W^{2} for the SDEs of SS and VV. One direction is to extend the scheme to the case without this assumption, i.e., the full constant correlation case, and analyse the convergence rate. This is nontrivial because the random variable NN and stochastic process rr in equation (1) then are not independent. The other direction is to extend the payoff PP to more complicated cases, such as those in Cozma, Mariapragassam and Reisinger [6].

Acknowledgements

This research is supported by National Natural Science Foundation of China (No.11801504).

References

  • [1] Alfonsi, A. (2005). On the discretization schemes for the CIR (and Bessel squared) processes. Monte Carlo Methods and Applications, 11(4), 355-384.
  • [2] Bakshi, S., Cao, C., Chen, Z. (2000). Pricing and hedging long-term options. Journal of Econometrics 94, 2003-2049.
  • [3] Black, F and Karasinski, P (1991). Bond and option pricing when short rates are lognormal, Financial Analysts Journal, 52-59.
  • [4] Brigo, D and Mercurio, F (2006). Interest rate models–theory and practice: with smile, inflation and credit, Springer Verlag.
  • [5] Cox, J. Ingersoll, J. and Ross, S (1985). A theory of term structure of interest rates. Econometrica, 53(2), 385-407.
  • [6] Cozma, A., Mariapragassam, M and Reisinger, C. (2018). Convergence of an Euler scheme for a hybrid stochastic-local volatility model with stochastic rates in foreign exchange markets, SIAM Journal on Financial Mathematics, 9, 127-170.
  • [7] Dereich, S. Neuenkirch, A. and Szpruch, L (2012). An Euler-type method for the strong approximation of the Cox-Ingersoll-Ross process. Proceedings of the Royal Society A, 468, 1105-1115.
  • [8] Dufresne, D (2001). The integrated square-root process. Working Paper, University of Montreal. https://minerva-access.unimelb.edu.au/handle/11343/33693.
  • [9] Giles, M (2008). Multilevel Monte Carlo path simulation. Operations Research, 56(3), 607-617.
  • [10] Giles, M (2008). Improved multilevel Monte Carlo convergence using the Milstein scheme. Monte Carlo and Quasi-Monte Carlo Methods, 343-358.
  • [11] Glasserman, P (2003). Monte Carlo Methods in Financial Engineering. Springer Sciences and Business media, New York.
  • [12] Grzelak, L.A and Oosterlee, C.W (2011). On the Heston model with stochastic interest rates. SIAM Journal on Financial Mathematics, 2(1), 255-286.
  • [13] Heston, S (1993). A closed-form solution for options with stochastic volatility with applications to bond and currency options. Review of Financial Studies, 6(2), 327 - 343.
  • [14] Hull, J.C and White, A (1990). Pricing interest rate derivative securities. Review of Financial Studies, 3(4), 573-592.
  • [15] Hurd, T and Kuznetsov, A (2008). Explicit formulas for Laplace transforms of stochastic integrals. Markov Processes and Related Fields, 14, 277-290.
  • [16] Kloeden, P and Platen, E (1999). Numerical Solution of Stochastic Differential Equations, 3rd edition, Springer Verlag, New York.
  • [17] Mickel, A. and Neuenkirch, A (2021). The weak convergence rate of two semi-exact discretization schemes for the Heston model. Risks, 9(1), 23.
  • [18] Neuenkirch, A. and Szpruch, L (2014). First order strong approximations of scalar SDEs with values in a domain. Numerische Mathematik, 128, 103-136.
  • [19] Protter, P (2005). Stochastic Integration and Differential Equations, 2nd edition, Springer.
  • [20] Rhee, C-H and Glynn, P.W (2015). Unbiased estimation with square root convergence for SDE models. Operations Research, 63(5), 1026-1043.
  • [21] Van Haastrecht, A and Pelsser, A (2011). Generic pricing of FX, inflation and stock options under stochastic interest rates and stochastic volatility, Quantitative Finance, 11, 665-691.
  • [22] Zheng, C (2017). Weak convergence rate of a time-discrete scheme for the Heston stochastic volatility model. SIAM Journal on Numerical Analysis, 55(3), 1243-1263.
  • [23] Zheng, C (2023). Multilevel Monte Carlo simulation for the Heston stochastic volatility model. Advances in Computational Mathematics, to appear.
  • [24] Zheng, C. Pan, J and Wang, Q (2023). Optimal distributions for randomized unbiased estimators with an infinite horizon and an adaptive algorithm. arXiv:2304.07797
  • [25] Zheng, Z., Blanchet, J. and Glynn, P (2018). Rates of convergence and CLTs for subcanonical debiased MLMC. In: Owen, A., Glynn, P. (eds) Monte Carlo and Quasi-Monte Carlo Methods in Scientific Computing 2016.