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

Positivity-preserving truncated Euler and Milstein methods for financial SDEs with super-linear coefficients

Shounian Deng Chen Fei Weiyin Fei wyfei@ahpu.edu.cn Xuerong Mao School of Mathematics-Physics and Finance, Anhui Polytechnic University, Wuhu 241000, China Business School, University of Shanghai for Science and Technology, Shanghai 200093, China Department of Mathematics and Statistics, University of Strathclyde, Glasgow G1 1XH, UK
Abstract

In this paper, we propose two variants of the positivity-preserving schemes, namely the truncated Euler-Maruyama (EM) method and the truncated Milstein scheme, applied to stochastic differential equations (SDEs) with positive solutions and super-linear coefficients. Under some regularity and integrability assumptions we derive the optimal strong convergence rates of the two schemes. Moreover, we demonstrate flexibility of our approaches by applying the truncated methods to approximate SDEs with super-linear coefficients (3/2 and Aït-Sahalia models) directly and also with sub-linear coefficients (CIR model) indirectly. Numerical experiments are provided to verify the effectiveness of the theoretical results.

keywords:
Truncated EM method, Truncated Milstein method, Strong convergence order, Positivity preservation.

1 Introduction

The goal of this paper is to derive a positivity-preserving numerical method for scalar SDEs which takes values in +\mathbb{R}_{+} and have super-linear coefficients. Typical examples of such SDEs in mathematical finance and bio-mathematical applications are the Heston-3/2 volatility process [1]

dX(t)=c1X(t)(c2X(t))dt+σX(t)3/2dB(t),X(0)=X0>0,\displaystyle dX(t)=c_{1}X(t)(c_{2}-X(t))dt+\sigma X(t)^{3/2}dB(t),\quad X(0)=X_{0}>0, (1.1)

with c1,c2,σ>0c_{1},c_{2},\sigma>0, and the Aït-Sahalia (AIT, for short) model [2]

dX(t)=\displaystyle dX(t)= (a1X(t)1a0+a1X(t)a2X(t)κ)dt+bX(t)θdB(t),X(0)=X0>0,\displaystyle\left(a_{-1}{X(t)^{-1}}-a_{0}+a_{1}X(t)-a_{2}X(t)^{\kappa}\right)dt+bX(t)^{\theta}dB(t),\quad X(0)=X_{0}>0, (1.2)

where all constant parameters are nonnegative, κ,θ>1\kappa,\theta>1, and B(t)B(t) is a Wiener process. The SDEs appearing in such models are highly nonlinear and may contain the singularity in the neighbourhood of zero in the drift or diffusion coefficient. Such SDEs in almost all cases cannot be solved explicitly, and it has been and still is a very active topic of research to approximate SDEs with super-linear coefficients; see, e.g., [3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20], and the references therein.

Table 1: Summary of the condition over the parameters for the convergence of the positivity-preserving schemes for AIT
Scheme Norm Parameter Range Rate
Lamperti Projected EM Lp,p1L^{p},p\geq 1 κ+1>2θ\kappa+1>2\theta 1p\frac{1}{p}
(Chassagneux et al. [5])
Lamperti Semi-Discrete EM L2L^{2} 12\frac{1}{2}
(Halidias and Stamatiou [21])
Lamperti BEM Lp,p2L^{p},p\geq 2 11
(Neuenkirch and Szpruch [4])
Logarithmic Truncated EM κ+3>4θ\kappa+3>4\theta 12\frac{1}{2}
(Yi et al. [11], Lei et al. [13])
Logarithmic Truncated EM
(Tang and Mao [22]) κ+1>2θ\kappa+1>2\theta
BEM without rate
(Szpruch et al. [3])
Theta EM L2L^{2} 12\frac{1}{2}
(Wang et al. [23])
Truncated EM Lp,p1L^{p},p\geq 1 12p\frac{1}{2p}
(Deng et al. [24])
Proposed Truncated EM 12\frac{1}{2}
Semi-implicit Tamed EM L2L^{2}
(Liu et al. [19])
Semi-implicit Projected Milstein 11
(Jiang et al. [18])
Proposed Truncated Milstein

Two significant challenges in constructing a numerical method for non-linear SDEs with positive solutions are to preserve positivity and to derive convergence with as high a rate as possible. In this section, we mainly discuss the positivity-preserving numerical methods for AIT (1.2). In fact, the techniques to handle the superlinearity and the singularity of the AIT coefficients can be applied to a more general SDE with positive solutions and super-linear coefficients. Table 1 gives a summary of the known results of some of the key methods which possess positivity preservation for AIT (1.2). We now discuss two main methods for the strong approximation of (1.2), the first is based on Lamperti/logarithmic transformation, and the second is direct approximation of AIT (1.2).

A classical transformation approach is to apply the Lamperti transformation Y=X1θY=X^{1-\theta} in order to obtain an auxiliary SDE in YY with a globally Lipschitz diffusion, but a drift function which is unbounded when solutions are in a neighbourhood of zero. For Lamperti transformation methods that preserve positivity for (1.2) see, for example, Lamperti backward EM (BEM) [4], Lamperti projected EM [5], and Lamperti Semi-Discrete EM [21]. However, this type of polynomial transformation would shift all the nonsmoothness into the drift function. In recent years, much effort has focused on deriving the combination of logarithmic transformation and Euler/Milstein-type schemes for general SDEs which have positive solutions. For such SDEs, an effective explicit scheme that preserves positivity is the logarithmic truncated EM proposed by Yi et al. [11]. The method is to apply a logarithmic transformation to original SDE and to numerically approximate the transformed SDE by a truncated EM (an explicit method established in [25, 26]). Under the Khasminskii-type and the global monotonicity conditions on coefficients of transformed SDE, the method was shown to have strong LpL^{p}-order of convergence 1/21/2. In a subsequent paper [16], a logarithmic truncated Milstein scheme was proved to have strong LpL^{p}-convergence of order 11, see Table 1. However, when the logarithmic method in [11] is applied to AIT (1.2), the method requires a more restrictive condition: κ>4θ3\kappa>4\theta-3 in order to satisfy the Khasminskii-type requirement for the transformed SDE, in contrast to the other positivity-preserving methods in Table 1. The same approach was also used by Tang and Mao in [22] to derive strong order 1/21/2 for the SDEs with positive solutions, but weaker assumptions on the coefficients of original SDE was required.

The second main approach that preserve positivity of the numerical solution is to directly discretise (1.2) by using the implicit or truncated tricks. A backward EM (BEM) was shown to preserve positivity of the solution for AIT (1.2) in Szpruch et al. [27], where the authors proved the strong convergence of the method, but without revealing a rate. This gap was filled by Wang et al. in [23], where the authors successfully recovered the mean-square convergence rate of order 1/21/2 for the theta EM applied to (1.2). However, the computational costs of solving implicit algebraic equations produced by classical BEM rise as the parameter κ\kappa increases.

A truncated EM that seeks to directly approximate AIT was proposed by Emmanuel and Mao in [28], where the authors proved the strong LpL^{p}-convergence of the method, but they did not reveal a convergence rate. Recently, the LpL^{p}-convergence rate of order 12p\frac{1}{2p} of this method was obtained by Deng et al. [24]. At the same time, a novel semi-implicit plus tamed EM was introduced by Liu et al. in [19], where the authors established the strong LpL^{p}-convergence of rate 1/21/2 based on an implicit treatment of the singular term a1x1a_{-1}x^{-1} to ensure the property of positiveness and a suitable taming of the super-linear terms a2xκ-a_{2}x^{\kappa} and bxθbx^{\theta}. It should be pointed out that this method is explicit, as the proposed scheme can be explicitly expressed by finding a positive root of a quadratic equation. Therefore, compared with the implicit schemes in [3, 4, 23] the computational costs in [19] can be significantly reduced. Similar approach was also used by Jiang et al. in [18] to derive strong order 11 of semi-implicit projected Milstein for AIT process. See Table 1 for a comparison.

The motivation of this paper is the following observation: in [24] the theoretical LpL^{p}-convergence rate of truncated EM for AIT is only 12p\frac{1}{2p}, rather than the optimal 12\frac{1}{2}, and it should be further increased. As a consequence, this paper is concerned with the optimal strong convergence rate of the positivity-preserving truncated methods for general SDE (2.2) with solution and super-linear coefficients. We note that the error analysis in [24] is based on the higher moment estimates of the numerical solution and the interpolation of the discretization scheme to continuous time. According to the theory of Mao [25, 26], it is inevitable to estimate the probability of truncated EM solution escaping from the truncated interval (1/R,R)(1/R,R), i.e., (ρΔ,RT)\mathbb{P}(\rho_{\Delta,R}\leq T). However, [24, Lemma 3.2] implies that the upper bound for this probability is of order 1/21/2 at most, i.e.,

(ρΔ,RT)CR(2γ+1)(γ+4)Δ1/2,\displaystyle\mathbb{P}(\rho_{\Delta,R}\leq T)\leq CR^{(2\gamma+1)\vee(\gamma+4)}\Delta^{1/2}, (1.3)

which prevents the truncated EM in [24] from achieving the optimal strong order 1/21/2. In this work, we seek to bypass this issue.

Motivated by the approaches of [5, 29], we adopt the error analysis based on the higher moment and inverse moment estimates of the true solution and the discretised scheme rather than the continuous time extension of numerical solution. We first prove a key Lemma 2.4 which shows the global monotonicity for the truncated functions and implies the stochastic C-Stability of the truncated EM in the sense of [29, Definition 3.2]. By the finiteness of the p0p_{0}th moments and p1p_{1}th inverse moments of the true solution XX, we then obtain the local truncations errors of f(X)f(X) and g(X)g(X) in LpL^{p}, which implies the stochastic B-Consistency of the truncated EM in the sense of [29, Definition 3.3], see Lemma 3.1. Combining Lemma 2.4 with Lemma 3.1 allows us to establish the optimal LpL^{p}-convergence of order 1/21/2 of positivity-preserving truncated EM for SDE (2.2) in the similar proofs as in [30, Theorem 5.36]. According to the global error estimate (3.14) in Theorem 3.2, we show that the truncated numerical solution YΔY_{\Delta} has some finite moments and inverse moments, see Proposition 3.4. By LpL^{p}-convergence Theorem 3.2 and moment boundedness Proposition 3.4, we apply truncated EM to directly approximate the 3/2 and the AIT models, see Applications 4.1 and 4.2. Based on the Lamperti transformation, we also apply this method to indirectly approximate the CIR process, see Application 4.3. Combining the Milstein scheme with the truncated strategy, we also prove the strong L2L^{2}-convergence order 11 for the truncated Milstein, in the similar method of analyzing the convergence rate of the truncated EM applied to SDE (2.2).

The main contributions of this paper are as follows:

  • 1.

    In contrast with the results of [24], our truncated EM for AIT obtains the optimal LpL^{p}-convergence of order 12\frac{1}{2}, which is far better than the rate 12p\frac{1}{2p} in [24].

  • 2.

    Compared with the projected EM in [5], where the method requires that the diffusion coefficients of SDEs are globally Lipschitz continuous, our truncated methods can apply to a wide class of SDEs with positive solutions and super-linearly growing diffusion coefficients.

  • 3.

    Our error analysis method does not relies on the high-order moments and inverse moments estimates, the continuous time extension of the numerical solution. Thus our truncated methods can be applied to approximate SDEs with super-linear coefficients directly and also with sub-linear coefficients indirectly, see Applications 4.

The structure of this paper is as follows. In Section 2 we give the form of the SDE with positive solution and super-linear coefficients, specify some regularity and integrability constraints placed upon the coefficients for the main strong convergence theorems. In Section 3 we set up the framework and prove the optimal strong convergence order as well as establish the moment boundedness results for the positivity-preserving truncated EM and Milstein schemes. In Section 4 we apply the proposed methods to some financial SDEs, such as the 3/2 and the AIT as well as the CIR models. In Section 5 we numerically compare convergence and efficiency of several commonly used methods. Finally, we have included in a small appendix a couple of proofs to make this paper self contained.

2 Setting and preliminaries

Let (Ω,,)(\Omega,{\mathcal{F}},\mathbb{P}) be a complete probability space with a filtration {t}t0\{{\cal F}_{t}\}_{t\geq 0} satisfying the usual conditions (i.e., it is increasing and right continuous while 0\cal{F}_{\textrm{0}} contains all \mathbb{P}-null sets). Let =(,)\mathbb{R}=(-\infty,\infty) and +=(0,)\mathbb{R}_{+}=(0,\infty). We write |||\cdot| and ,\langle\cdot,\cdot\rangle respectively for the Euclidean norm and the inner produce on \mathbb{R}. Let {B(t)}t0\{B(t)\}_{t\geq 0} be a one-dimensional Brownian motion with respect to the normal filtration {t}t0\{{\cal F}_{t}\}_{t\geq 0}. Let 𝔼\mathbb{E} denote the expectation. Denote by Lp(Ω;)L^{p}(\Omega;\mathbb{R}), for p>0p>0, the set of random variables ZZ such that Zp:=(𝔼[|Z|p])1/p<\|Z\|_{p}:=(\mathbb{E}[|Z|^{p}])^{1/p}<\infty. We denote by 𝔼t[X]:=𝔼[X|t]\mathbb{E}_{t}[X]:=\mathbb{E}[X|\mathcal{F}_{t}] the conditional expectation given the filtration t\mathcal{F}_{t}. For two real numbers aa and bb, ab:=max(a,b)a\vee b:=\max(a,b) and ab:=min(a,b)a\wedge b:=\min(a,b). W denote by 𝒞2(+)\mathcal{C}^{2}(\mathbb{R}_{+}) the space of twice differential functions with continuous derivatives on +\mathbb{R}_{+}. For any x+x\in\mathbb{R}_{+}, we denote x\lceil x\rceil the rounded up integer. For a set GG, its indicator function is denoted by 𝟏G\mathbf{1}_{G}. In the following it will be convenient to introduce the abbreviation

f(x):=df(x)dx,f′′(x):=d2f(x)dx2,andgg(x):=dg(x)dxg(x),x.\displaystyle f^{\prime}(x):=\frac{df(x)}{dx},\quad f^{\prime\prime}(x):=\frac{d^{2}f(x)}{dx^{2}},\quad\textrm{and}\quad g^{\prime}\cdot g(x):=\frac{dg(x)}{dx}g(x),\quad\forall x\in\mathbb{R}.

Moreover, given V(x)𝒞2(,+)V(x)\in\mathcal{C}^{2}(\mathbb{R},\mathbb{R}_{+}), we define a functional 𝕃V:\mathbb{L}V:\mathbb{R}\to\mathbb{R} by

𝕃V(x)=V(x)f(x)+12V′′(x)|g(x)|2,x.\displaystyle\mathbb{L}V(x)=V^{\prime}(x)f(x)+\frac{1}{2}V^{\prime\prime}(x)|g(x)|^{2},\quad\forall x\in\mathbb{R}.

In this work, we frequently apply the weighted Young inequality

xaybεxa+b+aa+b(aε(a+b))a/bya+b,x,y,a,b,ε>0.\displaystyle x^{a}y^{b}\leq\varepsilon x^{a+b}+\frac{a}{a+b}\left(\frac{a}{\varepsilon(a+b)}\right)^{a/b}y^{a+b},\quad\forall x,y,a,b,\varepsilon>0. (2.1)

Consider a one-dimensional SDE of the form

dX(t)\displaystyle dX(t) =f(X(t))dt+g(X(t))dB(t),t>0\displaystyle=f(X(t))dt+g(X(t))dB(t),\;t>0 (2.2)
X(0)\displaystyle X(0) =X0>0.\displaystyle=X_{0}>0.

Throughout this paper, we assume that SDE (2.2) has a unique strong solution in +\mathbb{R}_{+}. We now formulate the conditions on the drift and the diffusion coefficient functions.

Assumption 2.1

There are α0\alpha\geq 0, β0\beta\geq 0, K1>0K_{1}>0, such that

|f(x)f(y)||g(x)g(y)|K1(1+xα+yα+xβ+yβ)|xy|,x,y+.\displaystyle|f(x)-f(y)|\vee|g(x)-g(y)|\leq K_{1}\Big{(}1+x^{\alpha}+y^{\alpha}+x^{-\beta}+y^{-\beta}\Big{)}|x-y|,\quad\forall x,y\in\mathbb{R}_{+}. (2.3)
Assumption 2.2

Assume that f,g𝒞(+)f,g\in\mathcal{C}(\mathbb{R}_{+}) and there are positive numbers KK and q0>1q_{0}>1, such that

2f(x)+q0|g(x)|2K,x+.\displaystyle 2f^{\prime}(x)+q_{0}|g^{\prime}(x)|^{2}\leq K,\quad\forall x\in\mathbb{R}_{+}. (2.4)
Assumption 2.3

There are constants p0>2(α+1)p_{0}>2(\alpha+1) and p1>2βp_{1}>2\beta, such that

sup0tT𝔼[|X(t)|p0+|X(t)|p1]<.\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[|X(t)|^{p_{0}}+|X(t)|^{-p_{1}}\right]<\infty. (2.5)

In what follows, CC will be used to denote a positive constant depending on KK, TT, α\alpha, β\beta, p0p_{0}, p1p_{1} and X0X_{0}, but whose value may be different from line to line. We denote it by CpC_{p} if it depends on an extra parameter pp.

Assumption 2.2 implies that the monotonicity condition is fulfilled on +\mathbb{R}_{+}, i.e.,

2xy,f(x)f(y)+q0|g(x)g(y)|2K|xy|2,x,y+,\displaystyle 2\langle x-y,f(x)-f(y)\rangle+q_{0}|g(x)-g(y)|^{2}\leq K|x-y|^{2},\quad\forall x,y\in\mathbb{R}_{+}, (2.6)

and

|f(x)||g(x)|(12K1|f(1)|)(1+x1+α+xβ),x+,\displaystyle|f(x)|\vee|g(x)|\leq\Big{(}12K_{1}\vee|f(1)|\Big{)}\Big{(}1+x^{1+\alpha}+x^{-\beta}\Big{)},\quad\forall x\in\mathbb{R}_{+}, (2.7)

see [22, Remark 2.1]. Assumption 2.3 imposes a condition on the moments of the true solution XX in terms of the locally Lipschitz exponents α\alpha, β\beta. Combining this assumption and Assumption 2.1 allows us to bound the local truncations errors for f(X)f(X) and g(X)g(X), see Lemma 3.1.

Let Δ(0,1]\Delta\in(0,1], define a truncation mapping πΔ:[1R(Δ),R(Δ)]\pi_{\Delta}:\mathbb{R}\to\left[\frac{1}{R(\Delta)},R(\Delta)\right] by

πΔ(x)=1R(Δ)(xR(Δ)),x,\displaystyle\pi_{\Delta}(x)=\frac{1}{R(\Delta)}\vee(x\wedge R(\Delta)),\quad\forall x\in\mathbb{R}, (2.8)

where R(Δ):=L1ΔγR(\Delta):=L_{1}\Delta^{-\gamma} is the size of truncated interval with L1[1X0X0,)L_{1}\in\left[{\frac{1}{X_{0}}}\vee X_{0},\infty\right) and γ(0,12(αβ)]\gamma\in\left(0,\frac{1}{2(\alpha\vee\beta)}\right] is a constant to be determined later. Define the truncated functions

fΔ(x)=f(πΔ(x))andgΔ(x)=g(πΔ(x)),x.\displaystyle f_{\Delta}(x)=f(\pi_{\Delta}(x))\quad\textrm{and}\quad g_{\Delta}(x)=g(\pi_{\Delta}(x)),\quad\forall x\in\mathbb{R}. (2.9)

The following lemma shows that the truncated mapping πΔ\pi_{\Delta} is 11-Lipschitz continuous on \mathbb{R}, and the truncated functions fΔf_{\Delta} and gΔg_{\Delta} preserve the monotonicity condition (2.6).

Lemma 2.4

Let Assumption 2.2 hold. Let πΔ\pi_{\Delta}, fΔf_{\Delta} and gΔg_{\Delta} be the truncated mapping and functions defined in (2.8), (2.9) respectively. Then

|πΔ(x)πΔ(y)|\displaystyle|\pi_{\Delta}(x)-\pi_{\Delta}(y)| |xy|,x,y,\displaystyle\leq|x-y|,\quad\forall x,y\in\mathbb{R}, (2.10)
2xy,fΔ(x)fΔ(y)+\displaystyle 2\langle x-y,f_{\Delta}(x)-f_{\Delta}(y)\rangle+ q0|gΔ(x)gΔ(y)|22K|xy|2,x,y.\displaystyle q_{0}|g_{\Delta}(x)-g_{\Delta}(y)|^{2}\leq 2K|x-y|^{2},\quad\forall x,y\in\mathbb{R}. (2.11)

The proof of this lemma is postponed to Appendix 6.1. From Lemma 2.4, we have,

|πΔ(x)\displaystyle|\pi_{\Delta}(x) |x|+Δ,x,\displaystyle\leq|x|+\Delta,\quad\forall x\in\mathbb{R}, (2.12)
2x,fΔ(x)+(q01)\displaystyle 2\langle x,f_{\Delta}(x)\rangle+(q_{0}-1) |gΔ(x)|2K2(1+|x|2),x.\displaystyle|g_{\Delta}(x)|^{2}\leq K_{2}(1+|x|^{2}),\quad\forall x\in\mathbb{R}. (2.13)

Moreover, by Assumption 2.1 and (2.8) as well as (2.10), we have

|fΔ(x)fΔ(y)||gΔ(x)gΔ(y)|φ(Δ)|xy|,x,y,\displaystyle|f_{\Delta}(x)-f_{\Delta}(y)|\vee|g_{\Delta}(x)-g_{\Delta}(y)|\leq\varphi(\Delta)|x-y|,\quad\forall x,y\in\mathbb{R}, (2.14)

where φ(Δ)=CΔγ(αβ)\varphi(\Delta)=C\Delta^{-\gamma(\alpha\vee\beta)} with

φ(Δ)2ΔC.\displaystyle\varphi(\Delta)^{2}\Delta\leq C. (2.15)

We now provide two lemmas, their proofs are postponed to Appendix 6.2 and 6.3. Lemma 2.5 reveals the Hölder continuity of the true solution to (2.2) with respect to the norm in Lp(Ω;)L^{p}(\Omega;\mathbb{R}), while Lemma 2.6 shows the Lp(Ω;)L^{p}(\Omega;\mathbb{R})-error due to truncating the true solution XX on [1/R,R][1/R,R].

Lemma 2.5

Let Assumptions 2.1, 2.2 and 2.3 hold. Let p(0,p01+αp1β]p\in\left(0,\frac{p_{0}}{1+\alpha}\wedge\frac{p_{1}}{\beta}\right]. Then,

𝔼[|X(t+Δ)X(t)|p]CΔp/2,t[0,T],\displaystyle\mathbb{E}\left[|X(t+\Delta)-X(t)|^{p}\right]\leq C\Delta^{p/2},\quad t\in[0,T], (2.16)

where CC is a positive constant independent of Δ\Delta.

Lemma 2.6

Under the same assumption as Lemma 2.5, it holds that

𝔼[|f(X(t))fΔ(X(t)|p]𝔼[|g(X(t))gΔ(X(t)|p]1ψ(Δ)=C(Δγ(p0pαp)+Δγ(p1pβ+p)),t[0,T],\displaystyle\mathbb{E}\left[|f(X(t))-f_{\Delta}(X(t)|^{p}\right]\vee\mathbb{E}\left[|g(X(t))-g_{\Delta}(X(t)|^{p}\right]\leq\frac{1}{\psi(\Delta)}=C\Big{(}\Delta^{\gamma(p_{0}-p\alpha-p)}+\Delta^{\gamma(p_{1}-p\beta+p)}\Big{)},\quad t\in[0,T], (2.17)
𝔼[|X(t)πΔ(X(t))|p](Cp0R(Δ)p0p+Cp1R(Δ)p1+p)=C(Δγ(p0p)+Δγ(p1+p)),t[0,T],\displaystyle\mathbb{E}\left[|X(t)-\pi_{\Delta}(X(t))|^{p}\right]\leq\left(\frac{C_{p_{0}}}{R(\Delta)^{p_{0}-p}}+\frac{C_{p_{1}}}{R(\Delta)^{p_{1}+p}}\right)=C\Big{(}\Delta^{\gamma(p_{0}-p)}+\Delta^{\gamma(p_{1}+p)}\Big{)},\quad t\in[0,T], (2.18)

where ψ(Δ)=C(Δγ(p1pβ+p)+Δγ(p0pαp))\psi(\Delta)=C\Big{(}\Delta^{-\gamma(p_{1}-p\beta+p)}+\Delta^{-\gamma(p_{0}-p\alpha-p)}\Big{)} with p0p_{0}, p1p_{1} given by Assumption 2.3.

3 Main Results

3.1 Positivity-preserving Truncated EM (TEM) scheme

The first positivity-preserving scheme is termed truncated EM scheme. To be more precise, let Δ(0,1]\Delta\in(0,1] be a step size, define the following TEM scheme by setting XΔ(t0)=X0X_{\Delta}(t_{0})=X_{0} and computing

YΔ(tk)\displaystyle Y_{\Delta}(t_{k}) =πΔ(XΔ(tk)),\displaystyle=\pi_{\Delta}(X_{\Delta}(t_{k})),
XΔ(tk+1)\displaystyle X_{\Delta}(t_{k+1}) =XΔ(tk)+f(YΔ(tk))+g(YΔ(tk))ΔBk,\displaystyle=X_{\Delta}(t_{k})+f(Y_{\Delta}(t_{k}))+g(Y_{\Delta}(t_{k}))\Delta B_{k}, (3.1)

for k=0,1,2,k=0,1,2,\cdots, where ΔBk=B(tk+1)B(tk)\Delta B_{k}=B(t_{k+1})-B(t_{k}), tk=kΔt_{k}=k\Delta and πΔ\pi_{\Delta} has been defined in (2.8). By (2.9), (3.1) can be rewritten as the following form:

XΔ(tk+1)\displaystyle X_{\Delta}(t_{k+1}) =XΔ(tk)+fΔ(XΔ(tk))+gΔ(XΔ(tk))ΔBk,k=0,1,2,.\displaystyle=X_{\Delta}(t_{k})+f_{\Delta}(X_{\Delta}(t_{k}))+g_{\Delta}(X_{\Delta}(t_{k}))\Delta B_{k},\quad k=0,1,2,\cdots. (3.2)

We now begin to bound the local truncations errors for f(X)f(X) and g(X)g(X) in Lp(Ω;)L^{p}(\Omega;\mathbb{R}) sense.

Lemma 3.1

Let Assumptions 2.1, 2.2 and 2.3 hold with

p0(4α+2)andp14β.\displaystyle p_{0}\geq(4\alpha+2)\quad\textrm{and}\quad p_{1}\geq 4\beta. (3.3)

Let p[2,p02α+1p12β]p\in\left[2,\frac{p_{0}}{2\alpha+1}\wedge\frac{p_{1}}{2\beta}\right]. Then

𝔼[|Mk(d)|p]\displaystyle\mathbb{E}\left[|M^{(d)}_{k}|^{p}\right] CΔp(1ψ(Δ)+Δp/2),k=0,1,,\displaystyle\leq C\Delta^{p}\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)},\quad k=0,1,\cdots, (3.4)
𝔼[|Mk(w)|p]\displaystyle\mathbb{E}\left[|M^{(w)}_{k}|^{p}\right] CΔp/2(1ψ(Δ)+Δp/2),k=0,1,,\displaystyle\leq C\Delta^{p/2}\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)},\quad k=0,1,\cdots, (3.5)

where

Mk(d):=tktk+1(f(X(s))fΔ(X(tk)))𝑑s,\displaystyle M^{(d)}_{k}:=\int_{t_{k}}^{t_{k+1}}\Big{(}f(X(s))-f_{\Delta}(X(t_{k}))\Big{)}ds,
Mk(w):=tktk+1(g(X(s))gΔ(X(tk)))𝑑B(s),\displaystyle M^{(w)}_{k}:=\int_{t_{k}}^{t_{k+1}}\Big{(}g(X(s))-g_{\Delta}(X(t_{k}))\Big{)}dB(s),

and ψ\psi is defined (2.17).

Proof. One notes that condition (3.10) implies

pp0pα1α+1andpp1pβ1β.\displaystyle\frac{p}{p_{0}-p\alpha}\leq\frac{1}{\alpha+1}\quad\textrm{and}\quad\frac{p}{p_{1}-p\beta}\leq\frac{1}{\beta}. (3.6)

For any tk<stk+1t_{k}<s\leq t_{k+1}, we conclude from (2.3) and the Hölder inequality that

𝔼|f(X(s))f(X(tk))|p𝔼|g(X(s))g(X(tk))|p\displaystyle\mathbb{E}|f(X(s))-f(X(t_{k}))|^{p}\vee\mathbb{E}|g(X(s))-g(X(t_{k}))|^{p} (3.7)
C𝔼[(1+|X(s)|pα+|X(tk)|pα+|X(s)|pβ+|X(tk)|pβ)|X(s)X(tk)|p]\displaystyle\leq C\mathbb{E}\left[(1+|X(s)|^{p\alpha}+|X(t_{k})|^{p\alpha}+|X(s)|^{-p\beta}+|X(t_{k})|^{-p\beta})|X(s)-X(t_{k})|^{p}\right]
C(1+(𝔼|X(s)|p0)pα/p0+(𝔼|X(tk)|p0)pα/p0)(𝔼|X(s)X(tk)|pp0/(p0pα))(p0pα)/p0\displaystyle\leq C\left(1+\ (\mathbb{E}|X(s)|^{p_{0}})^{p\alpha/p_{0}}+(\mathbb{E}|X(t_{k})|^{p_{0}})^{p\alpha/p_{0}}\right)\left(\mathbb{E}|X(s)-X(t_{k})|^{pp_{0}/(p_{0}-p\alpha)}\right)^{(p_{0}-p\alpha)/p_{0}}
+C((𝔼|X(s)|p1)pβ/p1+(𝔼|X(tk)|p1)pβ/p1)(𝔼|X(s)X(tk)|pp1/(p1pβ))(p1pβ)/p1\displaystyle\quad+C\left(\ (\mathbb{E}|X(s)|^{-p_{1}})^{p\beta/p_{1}}+(\mathbb{E}|X(t_{k})|^{-p_{1}})^{p\beta/p_{1}}\right)\left(\mathbb{E}|X(s)-X(t_{k})|^{pp_{1}/(p_{1}-p\beta)}\right)^{(p_{1}-p\beta)/p_{1}}
CΔp/2,\displaystyle\leq C\Delta^{p/2},

where Assumption 2.3 and Lemma 2.5 have been used. Moreover,

Mk(d)=tktk+1(f(X(s))f(X(tk)))𝑑s+tktk+1(f(X(tk))fΔ(X(tk)))𝑑s.\displaystyle M^{(d)}_{k}=\int_{t_{k}}^{t_{k+1}}\Big{(}f(X(s))-f(X(t_{k}))\Big{)}ds+\int_{t_{k}}^{t_{k+1}}\Big{(}f(X(t_{k}))-f_{\Delta}(X(t_{k}))\Big{)}ds.

By the Hölder inequality, (2.17) and (3.7), we have

𝔼[|Mk(d)|p]\displaystyle\mathbb{E}\left[|M^{(d)}_{k}|^{p}\right] (3.8)
CΔp1tktk+1𝔼|f(X(s))f(X(tk))|p𝑑s+CΔp1tktk+1𝔼|f(X(tk))fΔ(X(tk))|p𝑑s\displaystyle\leq C\Delta^{p-1}\int_{t_{k}}^{t_{k+1}}\mathbb{E}\left|f(X(s))-f(X(t_{k}))\right|^{p}ds+C\Delta^{p-1}\int_{t_{k}}^{t_{k+1}}\mathbb{E}\left|f(X(t_{k}))-f_{\Delta}(X(t_{k}))\right|^{p}ds
CΔp(1ψ(Δ)+Δp/2).\displaystyle\leq C\Delta^{p}\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}.

By the Burkholder-Davis-Gundy identity, we have

𝔼[|Mk(w)|p]\displaystyle\mathbb{E}\left[|M^{(w)}_{k}|^{p}\right] (3.9)
=𝔼|tktk+1(g(X(s))gΔ(X(tk)))𝑑B(s)|p\displaystyle=\mathbb{E}\Big{|}\int_{t_{k}}^{t_{k+1}}\Big{(}g(X(s))-g_{\Delta}(X(t_{k}))\Big{)}dB(s)\Big{|}^{p}
C𝔼|tktk+1(g(X(s))g(X(tk)))𝑑B(s)|p+C𝔼|tktk+1(g(X(tk))gΔ(X(tk)))𝑑B(s)|p\displaystyle\leq C\mathbb{E}\Big{|}\int_{t_{k}}^{t_{k+1}}\Big{(}g(X(s))-g(X(t_{k}))\Big{)}dB(s)\Big{|}^{p}+C\mathbb{E}\Big{|}\int_{t_{k}}^{t_{k+1}}\Big{(}g(X(t_{k}))-g_{\Delta}(X(t_{k}))\Big{)}dB(s)\Big{|}^{p}
CΔp/21tktk+1𝔼|g(X(s))g(X(tk))|p𝑑s+CΔp/21tktk+1𝔼|g(X(tk))gΔ(X(tk))|p𝑑s.\displaystyle\leq C\Delta^{p/2-1}\int_{t_{k}}^{t_{k+1}}\mathbb{E}|g(X(s))-g(X(t_{k}))|^{p}ds+C\Delta^{p/2-1}\int_{t_{k}}^{t_{k+1}}\mathbb{E}|g(X(t_{k}))-g_{\Delta}(X(t_{k}))|^{p}ds.

Combining (3.7) with (2.17), we conclude from (3.9) that

𝔼[|Mk(w)|p]\displaystyle\mathbb{E}\left[|M^{(w)}_{k}|^{p}\right] CΔp/2(1ψ(Δ)+Δp/2).\displaystyle\leq C\Delta^{p/2}\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}.

The proof is finished. \Box

We now consider the global discretization error between the true solution XX and the discretized process XΔX_{\Delta}. The following theorem reveals the optimal strong convergence rate of TEM in the sense of Lp(Ω;)L^{p}(\Omega;\mathbb{R}).

Theorem 3.2 (Convergence order of TEM)

Let Assumptions 2.1, 2.2 and 2.3 hold with

p02(αβ)+2α+2andp12(αβ)+2β.\displaystyle p_{0}\geq 2(\alpha\vee\beta)+2\alpha+2\quad\textrm{and}\quad p_{1}\geq 2(\alpha\vee\beta)+2\beta. (3.10)

Let

p[2,p0(2α+1)(α+β+1)p1(α+β)2β].\displaystyle\displaystyle p\in\left[2,\frac{p_{0}}{(2\alpha+1)\vee(\alpha+\beta+1)}\wedge\frac{p_{1}}{(\alpha+\beta)\vee 2\beta}\right]. (3.11)

Then, the truncated EM scheme (3.1) by setting

R(Δ)=L1Δ12(αβ)\displaystyle R(\Delta)=L_{1}\Delta^{-\frac{1}{2(\alpha\vee\beta)}} (3.12)

has the property that

max0kT/Δ𝔼[|X(tk)XΔ(tk)|p]CΔp/2,T>0,\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-X_{\Delta}(t_{k})|^{p}]\leq C\Delta^{p/2},\quad T>0, (3.13)
max0kT/Δ𝔼[|X(tk)πΔ(XΔ(tk))|p]CΔp/2,T>0,\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}]\leq C\Delta^{p/2},\quad T>0, (3.14)

where CC is a positive constant independent of Δ\Delta.

Proof. One notes that condition (3.10) implies conditions (2.15) and (3.3). We observe from (2.2) that

X(tk+1)=X(tk)+fΔ(X(tk))Δ+gΔ(X(tk))ΔBk+Mk(d)+Mk(w),k=0,1,,\displaystyle X(t_{k+1})=X(t_{k})+f_{\Delta}(X(t_{k}))\Delta+g_{\Delta}(X(t_{k}))\Delta B_{k}+M^{(d)}_{k}+M^{(w)}_{k},\;k=0,1,\cdots, (3.15)

where Mk(d)M^{(d)}_{k} and Mk(w)M^{(w)}_{k} are defined in Lemma 3.1. Thus, we conclude from (3.2) and (3.15) that

X(tk+1)XΔ(tk+1)=\displaystyle X(t_{k+1})-X_{\Delta}(t_{k+1})= [X(tk)XΔ(tk)]+[fΔ(X(tk))fΔ(XΔ(tk))]Δ\displaystyle[X(t_{k})-X_{\Delta}(t_{k})]+[f_{\Delta}(X(t_{k}))-f_{\Delta}(X_{\Delta}(t_{k}))]\Delta
+[gΔ(X(tk))gΔ(XΔ(tk))]ΔBk+Mk(d)+Mk(w),k=0,1,.\displaystyle+[g_{\Delta}(X(t_{k}))-g_{\Delta}(X_{\Delta}(t_{k}))]\Delta B_{k}+M^{(d)}_{k}+M^{(w)}_{k},\;k=0,1,\cdots. (3.16)

Denote by

Ak:=X(tk+1)XΔ(tk+1),Fk:=fΔ(X(tk))fΔ(XΔ(tk)),Gk:=gΔ(X(tk))gΔ(XΔ(tk)).\displaystyle A_{k}:=X(t_{k+1})-X_{\Delta}(t_{k+1}),\quad F_{k}:=f_{\Delta}(X(t_{k}))-f_{\Delta}(X_{\Delta}(t_{k})),\quad G_{k}:=g_{\Delta}(X(t_{k}))-g_{\Delta}(X_{\Delta}(t_{k})). (3.17)

Without loss of generality, we assume that p2p\geq 2 is an even integer and Condition (3.11) is satisfied. It follows from the binomial formula that

Ak+1p\displaystyle A_{k+1}^{p} =(Ak+FkΔ+GkΔBk+Mk(d)+Mk(w))p=Akp+i=1p/2Π2i1,\displaystyle=(A_{k}+F_{k}\Delta+G_{k}\Delta B_{k}+M^{(d)}_{k}+M^{(w)}_{k})^{p}=A_{k}^{p}+\sum_{i=1}^{p/2}\Pi_{2i-1}, (3.18)

where

Πi\displaystyle\Pi_{i} =CpiAkpi(FkΔ+GkΔBk+Mk(d)+Mk(w))i,\displaystyle=C_{p}^{i}A_{k}^{p-i}(F_{k}\Delta+G_{k}\Delta B_{k}+M^{(d)}_{k}+M^{(w)}_{k})^{i},
+Cpi+1Akp(i+1)(FkΔ+GkΔBk+Mk(d)+Mk(w))i+1,i=1,3,,p/21.\displaystyle\quad+C_{p}^{i+1}A_{k}^{p-(i+1)}(F_{k}\Delta+G_{k}\Delta B_{k}+M^{(d)}_{k}+M^{(w)}_{k})^{i+1},\quad i=1,3,\cdots,p/2-1. (3.19)

By the elementary inequality (2.1), we have

Π1\displaystyle\Pi_{1} =pAkp1(FkΔ+GkΔBk+Mk(d)+Mk(w))\displaystyle=pA_{k}^{p-1}(F_{k}\Delta+G_{k}\Delta B_{k}+M^{(d)}_{k}+M^{(w)}_{k})
+p(p1)2Akp2(FkΔ+GkΔBk+Mk(d)+Mk(w))2\displaystyle\quad+\frac{p(p-1)}{2}A_{k}^{p-2}(F_{k}\Delta+G_{k}\Delta B_{k}+M^{(d)}_{k}+M^{(w)}_{k})^{2}
p2|Ak|p2(2AkFkΔ+q0|Gk|2|ΔBk|2)=:J1+C|Ak|p1|Mk(d)|=:J2+C|Ak|p2|Mk(w)|2=:J3\displaystyle\leq\underbrace{\frac{p}{2}|A_{k}|^{p-2}(2A_{k}F_{k}\Delta+q_{0}|G_{k}|^{2}|\Delta B_{k}|^{2})}_{=:J_{1}}+\underbrace{C|A_{k}|^{p-1}|M^{(d)}_{k}|}_{=:J_{2}}+\underbrace{C|A_{k}|^{p-2}|M^{(w)}_{k}|^{2}}_{=:J_{3}}
+C|Ak|p2|Fk|2Δ2=:J4+C|Ak|p2|Mk(d)|2=:J5+2Akp1GkΔBk+pAkp1Mk(w)=:J6\displaystyle\quad+\underbrace{C|A_{k}|^{p-2}|F_{k}|^{2}\Delta^{2}}_{=:J_{4}}+\underbrace{C|A_{k}|^{p-2}|M^{(d)}_{k}|^{2}}_{=:J_{5}}+\underbrace{2A_{k}^{p-1}G_{k}\Delta B_{k}+pA_{k}^{p-1}M^{(w)}_{k}}_{=:J_{6}}
=i=16Ji.\displaystyle=\sum_{i=1}^{6}J_{i}.

According to Lemma 2.4, we have

𝔼[J1]\displaystyle\mathbb{E}[J_{1}] =p2𝔼[|X(tk)XΔ(tk)|p2\displaystyle=\frac{p}{2}\mathbb{E}\Big{[}|X(t_{k})-X_{\Delta}(t_{k})|^{p-2}
×(2[X(tk)XΔ(tk)][fΔ(X(tk))fΔ(XΔ(tk))]Δ+q0[gΔ(X(tk))gΔ(XΔ(tk))]2|ΔBk|2)]\displaystyle\quad\times\Big{(}2[X(t_{k})-X_{\Delta}(t_{k})][f_{\Delta}(X(t_{k}))-f_{\Delta}(X_{\Delta}(t_{k}))]\Delta+q_{0}[g_{\Delta}(X(t_{k}))-g_{\Delta}(X_{\Delta}(t_{k}))]^{2}|\Delta B_{k}|^{2}\Big{)}\Big{]}
CΔ𝔼[|X(tk)XΔ(tk)|p].\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]. (3.20)

By the elementary inequality (2.1) and Lemma 3.1, we have

𝔼[J2]=C𝔼[|X(tk)XΔ(tk)|p1|Mk(d)|]\displaystyle\mathbb{E}[J_{2}]=C\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p-1}|M^{(d)}_{k}|\right] CΔ𝔼[|X(tk)XΔ(tk)|p]+C𝔼[|Mk(d)|p]Δp1\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\frac{\mathbb{E}\left[|M^{(d)}_{k}|^{p}\right]}{\Delta^{p-1}}
CΔ𝔼[|X(tk)XΔ(tk)|p]+CΔ(1ψ(Δ)+Δp/2).\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\Delta\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}. (3.21)

Similarly,

𝔼[J3]=C𝔼[|X(tk)XΔ(tk)|p2|Mk(w)|2]\displaystyle\mathbb{E}[J_{3}]=C\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p-2}|M^{(w)}_{k}|^{2}\right] CΔ𝔼[|X(tk)XΔ(tk)|p]+C𝔼[|Mk(w)|p]Δp/21\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\frac{\mathbb{E}\left[|M^{(w)}_{k}|^{p}\right]}{\Delta^{p/2-1}}
CΔ𝔼[|X(tk)XΔ(tk)|p]+CΔ(1ψ(Δ)+Δp/2),\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\Delta\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}, (3.22)

and

𝔼[J5]=C𝔼[|X(tk)XΔ(tk)|p2|Mk(d)|2]\displaystyle\mathbb{E}[J_{5}]=C\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p-2}|M^{(d)}_{k}|^{2}\right] CΔ𝔼[|X(tk)XΔ(tk)|p]+C𝔼[|Mk(d)|p]Δp/21\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\frac{\mathbb{E}\left[|M^{(d)}_{k}|^{p}\right]}{\Delta^{p/2-1}}
CΔ𝔼[|X(tk)XΔ(tk)|p]+CΔp/2+1(1ψ(Δ)+Δp/2)\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\Delta^{p/2+1}\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}
CΔ𝔼[|X(tk)XΔ(tk)|p]+CΔ(1ψ(Δ)+Δp/2).\displaystyle\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\Delta\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}. (3.23)

Moreover, fΔf_{\Delta} is globally Lipschitz continuous with Lipschitz constant φ(Δ)\varphi(\Delta), see (2.14). Thus

𝔼[J4]\displaystyle\mathbb{E}[J_{4}] =CΔ2𝔼[|X(tk)XΔ(tk)|p2|fΔ(X(tk))fΔ(XΔ(tk))|2]\displaystyle=C\Delta^{2}\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p-2}|f_{\Delta}(X(t_{k}))-f_{\Delta}(X_{\Delta}(t_{k}))|^{2}\right]
Cφ(Δ)2Δ2𝔼[|X(tk)XΔ(tk)|p]CΔ𝔼[|X(tk)XΔ(tk)|p].\displaystyle\leq C\varphi(\Delta)^{2}\Delta^{2}\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]. (3.24)

In addition, we have the following identity

𝔼[J6]\displaystyle\mathbb{E}[J_{6}] =p𝔼[|X(tk)XΔ(tk)|p1[gΔ(X(tk))gΔ(XΔ(tk))]ΔBk]\displaystyle=p\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p-1}\big{[}g_{\Delta}(X(t_{k}))-g_{\Delta}(X_{\Delta}(t_{k}))\big{]}\Delta B_{k}\right]
+p𝔼[(X(tk)XΔ(tk)|p1tktk+1(g(X(s))gΔ(X(tk)))dB(s)]=0.\displaystyle+p\mathbb{E}\left[(X(t_{k})-X_{\Delta}(t_{k})|^{p-1}\int_{t_{k}}^{t_{k+1}}\Big{(}g(X(s))-g_{\Delta}(X(t_{k}))\Big{)}dB(s)\right]=0. (3.25)

From (3.1)-(3.1), we have

𝔼[Π1]CΔ𝔼[|X(tk)XΔ(tk)|p]+CΔ(1ψ(Δ)+Δp/2).\displaystyle\mathbb{E}[\Pi_{1}]\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\Delta\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}. (3.26)

In the same fashion as (3.26) is obtained, we also can show that

i=2p/2𝔼[Π2i1]CΔ𝔼[|X(tk)XΔ(tk)|p]+CΔ(1ψ(Δ)+Δp/2).\displaystyle\sum_{i=2}^{p/2}\mathbb{E}\left[\Pi_{2i-1}\right]\leq C\Delta\mathbb{E}\left[|X(t_{k})-X_{\Delta}(t_{k})|^{p}\right]+C\Delta\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}. (3.27)

Thus, we conclude from (3.18) that

𝔼[|X(tk+1)XΔ(tk+1)|p]\displaystyle\mathbb{E}[|X(t_{k+1})-X_{\Delta}(t_{k+1})|^{p}] (3.28)
(1+CΔ)𝔼[|X(tk)XΔ(tk)|p]+CΔ(1ψ(Δ)+Δp/2),k=0,1,.\displaystyle\leq(1+C\Delta)\mathbb{E}[|X(t_{k})-X_{\Delta}(t_{k})|^{p}]+C\Delta\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)},\;k=0,1,\cdots.

Using the discrete Gronwall inequality yields

max0kN𝔼[|X(tk)XΔ(tk)|p]\displaystyle\max_{0\leq k\leq N}\mathbb{E}[|X(t_{k})-X_{\Delta}(t_{k})|^{p}] C(1ψ(Δ)+Δp/2)\displaystyle\leq C\Big{(}\frac{1}{\psi(\Delta)}+\Delta^{p/2}\Big{)}
=C(Δγ(p0pαp)+Δγ(p1pβ+p)+Δp/2).\displaystyle=C\Big{(}\Delta^{\gamma(p_{0}-p\alpha-p)}+\Delta^{\gamma(p_{1}-p\beta+p)}+\Delta^{p/2}\Big{)}. (3.29)

To balance the error terms, we set γ=12(αβ)\displaystyle\gamma=\frac{1}{2(\alpha\vee\beta)}. From this and (3.11), we have

γ(p0pαp)p/2andγ(p1pβ+p)p/2,\displaystyle\gamma(p_{0}-p\alpha-p)\geq p/2\quad\textrm{and}\quad\gamma(p_{1}-p\beta+p)\geq p/2, (3.30)

which implies that

sup0tT𝔼[|f(X(t))fΔ(X(t)|p|g(X(t))gΔ(X(t)|p|X(t)πΔ(X(t))|p]CΔp/2,\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[|f(X(t))-f_{\Delta}(X(t)|^{p}\vee|g(X(t))-g_{\Delta}(X(t)|^{p}\vee|X(t)-\pi_{\Delta}(X(t))|^{p}\right]\leq C\Delta^{p/2}, (3.31)

and (3.13) holds. Moreover, by Lemma 2.4 and (2.18) as well as (3.13), we have

𝔼[|X(tk)πΔ(XΔ(tk))|p]\displaystyle\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}] (3.32)
C𝔼[|X(tk)πΔ(X(tk))|p]+C𝔼[|πΔ(X(tk))πΔ(XΔ(tk))|p]\displaystyle\leq C\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X(t_{k}))|^{p}]+C\mathbb{E}[|\pi_{\Delta}(X(t_{k}))-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}]
C𝔼[|X(tk)πΔ(X(tk))|p]+C𝔼[|X(tk)XΔ(tk)|p]\displaystyle\leq C\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X(t_{k}))|^{p}]+C\mathbb{E}[|X(t_{k})-X_{\Delta}(t_{k})|^{p}]
CΔγ(p02)+CΔγ(p1+2)+CΔp/2CΔp/2,k=0,1,,\displaystyle\leq C\Delta^{\gamma(p_{0}-2)}+C\Delta^{\gamma(p_{1}+2)}+C\Delta^{p/2}\leq C\Delta^{p/2},\;k=0,1,\cdots,

which is the desired assertion (3.14). \Box

Remark 3.3

Let γ(0,12(αβ)]\gamma\in(0,\frac{1}{2(\alpha\vee\beta)}]. From the proofs of Theorem 3.2, we note that if Assumptions 2.1, 2.2 and 2.3 hold with p0(4α+2)p_{0}\geq(4\alpha+2) and p14βp_{1}\geq 4\beta, then for p[2,p02α+1p12β]p\in\left[2,\frac{p_{0}}{2\alpha+1}\wedge\frac{p_{1}}{2\beta}\right],

max0kT/Δ𝔼[|X(tk)πΔ(XΔ(tk))|p]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}] C(Δγ(p0pαp)+Δγ(p1pβ+p)+Δp/2).\displaystyle\leq C\Big{(}\Delta^{\gamma(p_{0}-p\alpha-p)}+\Delta^{\gamma(p_{1}-p\beta+p)}+\Delta^{p/2}\Big{)}. (3.33)

Comparing (3.33) with [31, Eq. 3.15], we observe that when 1/x1/x term varnishes in (2.2), then Theorem 3.2 degenerates to [31, Theorem 3.6].

For application use, we show that our numerical solutions have some finite, or inverse, moments.

Proposition 3.4 (Moment boundedness of TEM)

Let Assumptions 2.1, 2.2 and 2.3 hold with

p04(αβ)+8andp14(αβ)+2.\displaystyle p_{0}\geq 4(\alpha\vee\beta)+8\quad\textrm{and}\quad p_{1}\geq 4(\alpha\vee\beta)+2. (3.34)

Let

p[2,p02(αβ)+4p12(αβ)+1]\displaystyle\displaystyle p\in\left[2,\frac{p_{0}}{2(\alpha\vee\beta)+4}\wedge\frac{p_{1}}{2(\alpha\vee\beta)+1}\right] (3.35)

and R(Δ)=L1ΔγR(\Delta)=L_{1}\Delta^{-\gamma} with

γ=12(αβ)+4.\displaystyle\gamma=\frac{1}{2(\alpha\vee\beta)+4}. (3.36)

Then,

max0kT/Δ𝔼[|X(tk)πΔ(XΔ(tk))|p]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}] CΔp/2,T>0.\displaystyle\leq C\Delta^{p/2},\quad T>0. (3.37)

Moreover,

  1. (1).

    For 2qp(αβ)+2p+12\leq q\leq p(\alpha\vee\beta)+2p+1,

    max0kT/Δ𝔼[|πΔ(XΔ(tk))|q]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}] <,T>0;\displaystyle<\infty,\quad T>0; (3.38)
  2. (2).

    For 1qp(αβ)+11\leq q\leq p(\alpha\vee\beta)+1,

    max0kT/Δ𝔼[|πΔ(XΔ(tk))|q]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q}] <,T>0.\displaystyle<\infty,\quad T>0. (3.39)

Proof. Let pp and γ\gamma be the parameters satisfying (3.35) and (3.36), respectively. Then,

γ(p0pαp)\displaystyle\gamma(p_{0}-p\alpha-p) p[2(αβ)+4α1]2(αβ)+4[(αβ)+3]p2(αβ)+4p2,\displaystyle\geq\frac{p[2(\alpha\vee\beta)+4-\alpha-1]}{2(\alpha\vee\beta)+4}\geq\frac{[(\alpha\vee\beta)+3]p}{2(\alpha\vee\beta)+4}\geq\frac{p}{2},
γ(p1pβ+1)\displaystyle\gamma(p_{1}-p\beta+1) p[2(αβ)+1β+1]2(αβ)+4[(αβ)+2]p2(αβ)+4=p2.\displaystyle\geq\frac{p[2(\alpha\vee\beta)+1-\beta+1]}{2(\alpha\vee\beta)+4}\geq\frac{[(\alpha\vee\beta)+2]p}{2(\alpha\vee\beta)+4}=\frac{p}{2}.

Therefore, (3.37) follows from (3.33). We now prove (1). For 2qp2\leq q\leq p, (3.38) follows from Theorem 3.2, Assumption 2.3 and the following inequality

|πΔ(XΔ(tk))|p2p1|X(tk)|p+2p1|X(tk)πΔ(XΔ(tk))|p.\displaystyle|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}\leq 2^{p-1}|X(t_{k})|^{p}+2^{p-1}|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}.

We now assume that p<qp0p<q\leq p_{0}. Write R=R(Δ)R=R(\Delta) for simplicity. Let

πΔ(XΔ(tk))=ekΔ+X(tk),\displaystyle\pi_{\Delta}(X_{\Delta}(t_{k}))=e^{\Delta}_{k}+X(t_{k}),

where ekΔ:=πΔ(XΔ(tk))X(tk)e^{\Delta}_{k}:=\pi_{\Delta}(X_{\Delta}(t_{k}))-X(t_{k}). Clearly, we have that

|πΔ(XΔ(tk))|q=\displaystyle|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}= |πΔ(XΔ(tk))|q𝟏{|X(tk)|>R}+|πΔ(XΔ(tk))|q𝟏{|X(tk)|R,|ekΔ|1}\displaystyle|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}\mathbf{1}_{\{|X(t_{k})|>R\}}+|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|\leq 1\}} (3.40)
+|πΔ(XΔ(tk))|q𝟏{|X(tk)|R,|ekΔ|>1}.\displaystyle+|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|>1\}}.

Since, 1/RπΔ(XΔ(tk))R1/R\leq\pi_{\Delta}(X_{\Delta}(t_{k}))\leq R, we conclude from Assumption 2.3 that

𝔼[|πΔ(XΔ(tk))|q𝟏{|X(tk)|>R}]\displaystyle\mathbb{E}\Big{[}|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}\mathbf{1}_{\{|X(t_{k})|>R\}}\Big{]}
Rq𝔼[𝟏{|X(tk)|>R}]Rq𝔼[|X(tk)|q]Rq𝔼[|X(tk)|p0]C.\displaystyle\leq R^{q}\mathbb{E}\Big{[}\mathbf{1}_{\{|X(t_{k})|>R\}}\Big{]}\leq R^{q}\frac{\mathbb{E}\left[|X(t_{k})|^{q}\right]}{R^{q}}\leq\mathbb{E}\left[|X(t_{k})|^{p_{0}}\right]\leq C. (3.41)

By the elementary inequality and Assumption 2.3, we have

𝔼[|πΔ(XΔ(tk))|q𝟏{|X(tk)|R,|ekΔ|1}]\displaystyle\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|\leq 1\}}\right] =𝔼[|X(tk)+ekΔ|q𝟏{|X(tk)|R,|ekΔ|1}]\displaystyle=\mathbb{E}\left[|X(t_{k})+e^{\Delta}_{k}|^{q}\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|\leq 1\}}\right] (3.42)
C(𝔼[|X(tk)|q]+𝔼[|ekΔ|p]𝟏{|X(tk)|R,|ekΔ|1})\displaystyle\leq C\Big{(}\mathbb{E}[|X(t_{k})|^{q}]+\mathbb{E}[|e^{\Delta}_{k}|^{p}]\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|\leq 1\}}\Big{)}
C(𝔼[|X(tk)|q]+1)C.\displaystyle\leq C\Big{(}\mathbb{E}[|X(t_{k})|^{q}]+1\Big{)}\leq C.

We note that

|y|q|x|q+Cq(|x|q1+|y|q1)|xy|,x,y.\displaystyle|y|^{q}\leq|x|^{q}+C_{q}(|x|^{q-1}+|y|^{q-1})|x-y|,\quad\forall x,y\in\mathbb{R}.

Therefore,

|πΔ(XΔ(tk))|q𝟏{|X(tk)|R,|ekΔ|>1}\displaystyle|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|>1\}} (3.43)
|X(tk)|q+Cq(|X(tk)|q1+|πΔ(XΔ(tk))|q1)|ekΔ|𝟏{|X(tk)|R,|ekΔ|>1}.\displaystyle\leq|X(t_{k})|^{q}+C_{q}\Big{(}|X(t_{k})|^{q-1}+|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q-1}\Big{)}|e^{\Delta}_{k}|\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|>1\}}.

By the Hölder and the Markov inequalities as well as Theorem 3.2, we have

𝔼[|ekΔ|𝟏{|ekΔ|>1}]\displaystyle\mathbb{E}\left[|e^{\Delta}_{k}|\mathbf{1}_{\{|e^{\Delta}_{k}|>1\}}\right] (𝔼[|ekΔ|p])1/p(𝔼[𝟏{|ekΔ|>1}])11/p\displaystyle\leq\Big{(}\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}\Big{)}^{1/p}\Big{(}\mathbb{E}\left[\mathbf{1}_{\{|e^{\Delta}_{k}|>1\}}\right]\Big{)}^{1-1/p} (3.45)
(𝔼[|ekΔ|p])1/p(𝔼[|ekΔ|p])11/p=𝔼[|ekΔ|p].\displaystyle\leq\Big{(}\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}\Big{)}^{1/p}\Big{(}\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}\Big{)}^{1-1/p}=\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}.

If we set

qp2γ+1=p[(αβ)+2]+1\displaystyle q\leq\frac{p}{2\gamma}+1=p[(\alpha\vee\beta)+2]+1 (3.46)

such that γ(q1)p/2\gamma(q-1)\leq p/2, then we conclude from (3.43) and (3.45) that

𝔼[|πΔ(XΔ(tk))|q𝟏{|X(tk)|R,|ekΔ|>1}]\displaystyle\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|>1\}}\right] (3.47)
𝔼[|X(tk)|q]+CRq1𝔼[|ekΔ|𝟏{|X(tk)|R,|ekΔ|>1}]\displaystyle\leq\mathbb{E}\Big{[}|X(t_{k})|^{q}\Big{]}+CR^{q-1}\mathbb{E}\left[|e^{\Delta}_{k}|\mathbf{1}_{\{|X(t_{k})|\leq R,\;|e^{\Delta}_{k}|>1\}}\right]
C(1+Rq1)𝔼[|ekΔ|𝟏{|ekΔ|>1}]\displaystyle\leq C(1+R^{q-1})\mathbb{E}\left[|e^{\Delta}_{k}|\mathbf{1}_{\{|e^{\Delta}_{k}|>1\}}\right]
CRq1𝔼[|ekΔ|p]CRq1Δp/2=CΔp/2γ(q1)C.\displaystyle\leq CR^{q-1}\mathbb{E}\left[|e^{\Delta}_{k}|^{p}\right]\leq CR^{q-1}\Delta^{p/2}=C\Delta^{p/2-\gamma(q-1)}\leq C.

From (3.40)-(3.47), we conclude that (3.38) holds.

We now prove (2). For q1q\geq 1 We observe that

|πΔ(XΔ(tk))|q\displaystyle|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q} (3.48)
=\displaystyle= |πΔ(XΔ(tk))|q𝟏{|X(tk)|1>R}+|πΔ(XΔ(tk))|q𝟏{|X(tk)|1R,|ekΔ|R2}\displaystyle|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q}\mathbf{1}_{\{|X(t_{k})|^{-1}>R\}}+|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q}\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|\leq R^{-2}\}}
+|πΔ(XΔ(tk))|q𝟏{|X(tk)|1R,|ekΔ|>R2}.\displaystyle+|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q}\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|>R^{-2}\}}.

Since, 1/R1/πΔ(XΔ(tk))R1/R\leq 1/\pi_{\Delta}(X_{\Delta}(t_{k}))\leq R, we conclude from Assumption 2.3 that

𝔼[|πΔ(XΔ(tk))|q𝟏{|X(tk)|1>R}]Rq𝔼[𝟏{|X(tk)|1>R}]\displaystyle\mathbb{E}\Big{[}|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q}\mathbf{1}_{\{|X(t_{k})|^{-1}>R\}}\Big{]}\leq R^{q}\mathbb{E}\Big{[}\mathbf{1}_{\{|X(t_{k})|^{-1}>R\}}\Big{]} (3.49)
Rq𝔼[|X(tk)|q]RqCq.\displaystyle\leq R^{q}\frac{\mathbb{E}\left[|X(t_{k})|^{-q}\right]}{R^{q}}\leq C_{q}.

By the elementary inequality and Assumption 2.3, we have

𝔼[|πΔ(XΔ(tk))|q𝟏{|X(tk)|1R,|ekΔ|R2}]\displaystyle\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q}\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|\leq R^{-2}\}}\right] (3.50)
C(𝔼[|1X(tk)|q]+𝔼[|1X(tk)1πΔ(X(tk))|q𝟏{|X(tk)|1R,|ekΔ|R2}])\displaystyle\leq C\left(\mathbb{E}\left[\Big{|}\frac{1}{X(t_{k})}\Big{|}^{q}\right]+\mathbb{E}\left[\Big{|}\frac{1}{X(t_{k})}-\frac{1}{\pi_{\Delta}(X(t_{k}))}\Big{|}^{q}\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|\leq R^{-2}\}}\right]\right)
C(𝔼[|1X(tk)|q]+𝔼[|ekΔ|q|X(tk)|q|πΔ(XΔ(tk))|q𝟏{|X(tk)|1R,|ekΔ|R2}])\displaystyle\leq C\left(\mathbb{E}\left[\Big{|}\frac{1}{X(t_{k})}\Big{|}^{q}\right]+\mathbb{E}\left[\frac{|e^{\Delta}_{k}|^{q}}{|X(t_{k})|^{q}|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{q}}\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|\leq R^{-2}\}}\right]\right)
C(𝔼[|1X(tk)|q]+R2q𝔼[|ekΔ|q𝟏{|X(tk)|1R,|ekΔ|R2}])\displaystyle\leq C\left(\mathbb{E}\left[\Big{|}\frac{1}{X(t_{k})}\Big{|}^{q}\right]+R^{2q}\mathbb{E}\left[|e^{\Delta}_{k}|^{q}\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|\leq R^{-2}\}}\right]\right)
C(𝔼[|1X(tk)|q]+1)C.\displaystyle\leq C\left(\mathbb{E}\left[\Big{|}\frac{1}{X(t_{k})}\Big{|}^{q}\right]+1\right)\leq C.

By the Hölder and Markov inequality, we have

𝔼[|ekΔ|𝟏{|ekΔ|>R2}]\displaystyle\mathbb{E}\left[|e^{\Delta}_{k}|\mathbf{1}_{\{|e^{\Delta}_{k}|>R^{-2}\}}\right] (𝔼[|ekΔ|p])1/p(𝔼[𝟏{|ekΔ|>R2}])11/p\displaystyle\leq\Big{(}\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}\Big{)}^{1/p}\Big{(}\mathbb{E}\left[\mathbf{1}_{\{|e^{\Delta}_{k}|>R^{-2}\}}\right]\Big{)}^{1-1/p}
(𝔼[|ekΔ|p])1/p(R2p𝔼[|ekΔ|p])11/p=R2p2𝔼[|ekΔ|p].\displaystyle\leq\Big{(}\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}\Big{)}^{1/p}\Big{(}R^{2p}\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}\Big{)}^{1-1/p}=R^{2p-2}\mathbb{E}\Big{[}|e^{\Delta}_{k}|^{p}\Big{]}. (3.51)

If we set

qp2γ2p+1=p(αβ)+1\displaystyle q\leq\frac{p}{2\gamma}-2p+1=p(\alpha\vee\beta)+1 (3.52)

such that γ(q+2p1)p/2\gamma(q+2p-1)\leq p/2, then

𝔼[|πΔ(XΔ(tk))|q𝟏{|X(tk)|1R,|ekΔ|>R2}]\displaystyle\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-q}\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|>R^{-2}\}}\right] (3.53)
𝔼[|X(tk)|q]+C𝔼[(|X(tk)|q1+|πΔ(X(tk))|p1)|ekΔ|𝟏{|X(tk)|1R,|ekΔ|>R2}]\displaystyle\leq\mathbb{E}\Big{[}|X(t_{k})|^{-q}\Big{]}+C\mathbb{E}\left[\Big{(}|X(t_{k})|^{-q-1}+|\pi_{\Delta}(X(t_{k}))|^{-p-1}\Big{)}|e^{\Delta}_{k}|\mathbf{1}_{\{|X(t_{k})|^{-1}\leq R,\;|e^{\Delta}_{k}|>R^{-2}\}}\right]
𝔼[|X(tk)|q]+CqRq+1𝔼[|ekΔ|𝟏{|X(tk)|qR,|ekΔ|>R2}]\displaystyle\leq\mathbb{E}\Big{[}|X(t_{k})|^{-q}\Big{]}+C_{q}R^{q+1}\mathbb{E}\left[|e^{\Delta}_{k}|\mathbf{1}_{\{|X(t_{k})|^{-q}\leq R,\;|e^{\Delta}_{k}|>R^{-2}\}}\right]
C(1+Rq+1)𝔼[|ekΔ|𝟏{|ekΔ|>R2}]\displaystyle\leq C(1+R^{q+1})\mathbb{E}\left[|e^{\Delta}_{k}|\mathbf{1}_{\{|e^{\Delta}_{k}|>R^{-2}\}}\right]
C(1+Rq+1)R2p2𝔼[|ekΔ|p]\displaystyle\leq C(1+R^{q+1})R^{2p-2}\mathbb{E}\left[|e^{\Delta}_{k}|^{p}\right]
CRq+2p1Δp/2CΔp/2γ(q+2p1)C.\displaystyle\leq CR^{q+2p-1}\Delta^{p/2}\leq C\Delta^{p/2-\gamma(q+2p-1)}\leq C.

From (3.48)-(3.53), we conclude that (3.39) holds. \Box

From the above analysis, we observe that in L2(Ω;)L^{2}(\Omega;\mathbb{R}) sense if we set γ=1(p02α2)(p12β+2)\gamma=\frac{1}{(p_{0}-2\alpha-2)\wedge(p_{1}-2\beta+2)}, then our convergence results can be described in a more concise form, see the following corollary.

Corollary 3.5

Let Assumptions 2.1, 2.2 and 2.3 hold with p0=p1p_{0}=p_{1}, α=β\alpha=\beta and p0=(4α+2)p_{0}=(4\alpha+2). Let R(Δ)=L1Δ12αR(\Delta)=L_{1}\Delta^{-\frac{1}{2\alpha}}. Then

max0kT/Δ𝔼[|X(tk)πΔ(XΔ(tk))|2]CΔ,T>0,\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{2}]\leq C\Delta,\quad T>0, (3.54)

and

max0kT/Δ𝔼[|πΔ(XΔ(tk))|p022]<T>0.\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{\frac{p_{0}}{2}\vee 2}\right]<\infty\quad T>0. (3.55)

Moreover, if p010p_{0}\geq 10, then

max0kT/Δ𝔼[|πΔ(XΔ(tk))|(p024)]<,T>0.\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-\left(\frac{p_{0}}{2}-4\right)}\right]<\infty,\quad T>0. (3.56)
Remark 3.6

It is worth mentioning how our results of TEM compares with that of Zhan and Li in [32], where the authors proved the L2L^{2}-convergence of order 1/21/2 of a truncated EM for super-linear SDE. However, their method do not preserve the property of positiveness that we are interested in and their results do not reveal the boundedness of moment and inverse moment of the numerical solutions. This hinders the further application of the truncated method for some important SDEs, such as Lamperti transformed CIR model, see Application 4.3.

3.2 Positivity-preserving Truncated Milstein (TMil) scheme

We now come to the second numerical scheme, which is called truncated Milstein (TMil) scheme. In order to show the first-order strong rate of convergence for this scheme, we need the following assumption, which is stronger than Assumption 2.1.

Assumption 3.7

Let f,g𝒞2(+)f,g\in\mathcal{C}^{2}(\mathbb{R}_{+}). There are α^1\hat{\alpha}\geq 1, β^0\hat{\beta}\geq 0, K^1>0\hat{K}_{1}>0, such that for any x,y+x,y\in\mathbb{R}_{+} it holds

|f′′(x)|\displaystyle|f^{\prime\prime}(x)| K^1(1+xα^1+xβ^),\displaystyle\leq\hat{K}_{1}\Big{(}1+x^{\hat{\alpha}-1}+x^{-\hat{\beta}}\Big{)}, (3.57)
|g′′(x)|\displaystyle|g^{\prime\prime}(x)| K^1(1+xα^/21+xβ^/2).\displaystyle\leq\hat{K}_{1}\Big{(}1+x^{\hat{\alpha}/2-1}+x^{-\hat{\beta}/2}\Big{)}. (3.58)

For a possibly enlarged CC the following estimates are an immediate consequence of Assumption 3.7 and the mean value theorem: For any x,y+x,y\in\mathbb{R}_{+} it holds

|f(x)f(y)|C(1+xα^1+yα^1+xβ^+yβ^)|xy|,\displaystyle|f^{\prime}(x)-f^{\prime}(y)|\leq C\Big{(}1+x^{\hat{\alpha}-1}+y^{\hat{\alpha}-1}+x^{-\hat{\beta}}+y^{-\hat{\beta}}\Big{)}|x-y|, (3.59)
|f(x)f(y)|C(1+xα^+yα^+xβ^+yβ^)|xy|,\displaystyle|f(x)-f(y)|\leq C\Big{(}1+x^{\hat{\alpha}}+y^{\hat{\alpha}}+x^{-\hat{\beta}}+y^{-\hat{\beta}}\Big{)}|x-y|, (3.60)
|g(x)g(y)|C(1+xα^/21+yα^/21+xβ^/2+yβ^/2)|xy|,\displaystyle|g^{\prime}(x)-g^{\prime}(y)|\leq C\Big{(}1+x^{\hat{\alpha}/2-1}+y^{\hat{\alpha}/2-1}+x^{-\hat{\beta}/2}+y^{-\hat{\beta}/2}\Big{)}|x-y|, (3.61)
|g(x)g(y)|C(1+xα^/2+yα^/2+xβ^/2+yβ^/2)|xy|,\displaystyle|g(x)-g(y)|\leq C\Big{(}1+x^{\hat{\alpha}/2}+y^{\hat{\alpha}/2}+x^{-\hat{\beta}/2}+y^{-\hat{\beta}/2}\Big{)}|x-y|, (3.62)

and

|f(x)|C(1+xα^+xβ^),\displaystyle|f^{\prime}(x)|\leq C\Big{(}1+x^{\hat{\alpha}}+x^{-\hat{\beta}}\Big{)}, (3.63)
|f(x)|C(1+xα^+1+xβ^),\displaystyle|f(x)|\leq C\Big{(}1+x^{\hat{\alpha}+1}+x^{-\hat{\beta}}\Big{)}, (3.64)
|g(x)|C(1+xα^/2+xβ^/2),\displaystyle|g^{\prime}(x)|\leq C\Big{(}1+x^{\hat{\alpha}/2}+x^{-\hat{\beta}/2}\Big{)}, (3.65)
|g(x)|C(1+xα^/2+1+xβ^/2),\displaystyle|g(x)|\leq C\Big{(}1+x^{\hat{\alpha}/2+1}+x^{-\hat{\beta}/2}\Big{)}, (3.66)

As above, we verify under Assumption 3.7 that the mapping ggg^{\prime}\cdot g satisfies the polynomial Lipschitz condition

|gg(x)gg(y)|C(1+xα^+yα^+xβ^+yβ^)|xy|,\displaystyle|g^{\prime}\cdot g(x)-g^{\prime}\cdot g(y)|\leq C\Big{(}1+x^{\hat{\alpha}}+y^{\hat{\alpha}}+x^{-\hat{\beta}}+y^{-\hat{\beta}}\Big{)}|x-y|, (3.67)

and the polynomial growth bound

|gg(x)|C(1+xα^+1+xβ^),\displaystyle|g^{\prime}\cdot g(x)|\leq C\Big{(}1+x^{\hat{\alpha}+1}+x^{-\hat{\beta}}\Big{)}, (3.68)

for any x,y+x,y\in\mathbb{R}_{+}. We define the following notation for the stochastic increments:

It,s:=tstt1𝑑B(t2)𝑑B(t1)=12([B(s)B(t)]2(st)),0t<s.\displaystyle I_{t,s}:=\int_{t}^{s}\int_{t}^{t_{1}}dB(t_{2})dB(t_{1})=\frac{1}{2}\Big{(}[B(s)-B(t)]^{2}-(s-t)\Big{)},\quad 0\leq t<s.

We now begin to construct a positivity-preserving truncated Milstein method, which is defined by setting X^Δ(t0)=X0\hat{X}_{\Delta}(t_{0})=X_{0} and by the recursion

Y^Δ(tk)\displaystyle\hat{Y}_{\Delta}(t_{k}) =πΔ(X^Δ(tk)),\displaystyle=\pi_{\Delta}(\hat{X}_{\Delta}(t_{k})),
X^Δ(tk+1)\displaystyle\hat{X}_{\Delta}(t_{k+1}) =X^Δ(tk)+f(Y^Δ(tk))+g(Y^Δ(tk))ΔBk+12gg(Y^Δ(tk))(|ΔBk|2Δ),\displaystyle=\hat{X}_{\Delta}(t_{k})+f(\hat{Y}_{\Delta}(t_{k}))+g(\hat{Y}_{\Delta}(t_{k}))\Delta B_{k}+\frac{1}{2}g^{\prime}\cdot g(\hat{Y}_{\Delta}(t_{k}))(|\Delta B_{k}|^{2}-\Delta), (3.69)
=X^Δ(tk)+fΔ(X^Δ(tk))+gΔ(X^Δ(tk))ΔBk+ggΔ(X^Δ(tk))Itk,tk+1,\displaystyle=\hat{X}_{\Delta}(t_{k})+f_{\Delta}(\hat{X}_{\Delta}(t_{k}))+g_{\Delta}(\hat{X}_{\Delta}(t_{k}))\Delta B_{k}+g^{\prime}\cdot g_{\Delta}(\hat{X}_{\Delta}(t_{k}))I_{t_{k},t_{k+1}}, (3.70)

for k=0,1,2,k=0,1,2,\cdots, where πΔ\pi_{\Delta} has been defined in (2.8) with a given parameter γ(0,12(α^β^)]\gamma\in\left(0,\frac{1}{2(\hat{\alpha}\vee\hat{\beta})}\right], fΔf_{\Delta} and gΔg_{\Delta} are defined in (2.9),

ggΔ(x):=gg(πΔ(x)),x.\displaystyle g^{\prime}\cdot g_{\Delta}(x):=g^{\prime}\cdot g(\pi_{\Delta}(x)),\quad\forall x\in\mathbb{R}. (3.71)

The following lemma implies that the TMil method is stochastically CC-stable in the sense of [29, Definition 3.2], and plays an important role in the convergence analysis of the TMil method.

Lemma 3.8

Consider mapping πΔ\pi_{\Delta} defined in (2.8) with γ(0,12(α^β^)]\gamma\in\left(0,\frac{1}{2(\hat{\alpha}\vee\hat{\beta})}\right]. Let Assumption 2.2 hold. Then there exists a constant KK only depending on q0q_{0} such that

|(xy)Δ(fΔ(x)fΔ(y))|+q0Δ|gΔ(x)gΔ(y)|2+q0Δ2|ggΔ(x)ggΔ(y)|2\displaystyle|(x-y)-\Delta(f_{\Delta}(x)-f_{\Delta}(y))|+q_{0}\Delta|g_{\Delta}(x)-g_{\Delta}(y)|^{2}+q_{0}\Delta^{2}|g^{\prime}\cdot g_{\Delta}(x)-g^{\prime}\cdot g_{\Delta}(y)|^{2}
(1+CΔ)|xy|2,x,y.\displaystyle\leq(1+C\Delta)|x-y|^{2},\quad\forall x,y\in\mathbb{R}.

Proof. By (3.67) and (2.10), we have

|ggΔ(x)ggΔ(y)|C(1+Rα^+Rβ^)|πΔ(x)πΔ(y)|CΔγ(α^β^)|xy|,x,y.\displaystyle|g^{\prime}\cdot g_{\Delta}(x)-g^{\prime}\cdot g_{\Delta}(y)|\leq C(1+R^{\hat{\alpha}}+R^{\hat{\beta}})|\pi_{\Delta}(x)-\pi_{\Delta}(y)|\leq C\Delta^{-\gamma(\hat{\alpha}\vee\hat{\beta})}|x-y|,\quad\forall x,y\in\mathbb{R}.

Thus, by Lemma 2.4 and (2.14), we have

|(xy)Δ(fΔ(x)fΔ(y))|+q0Δ|gΔ(x)gΔ(y)|2+q0Δ2|ggΔ(x)ggΔ(y)|2\displaystyle|(x-y)-\Delta(f_{\Delta}(x)-f_{\Delta}(y))|+q_{0}\Delta|g_{\Delta}(x)-g_{\Delta}(y)|^{2}+q_{0}\Delta^{2}|g^{\prime}\cdot g_{\Delta}(x)-g^{\prime}\cdot g_{\Delta}(y)|^{2}
=|xy|2+2Δxy,fΔ(x)fΔ(y)+q0Δ|gΔ(x)gΔ(y)|2\displaystyle=|x-y|^{2}+2\Delta\langle x-y,f_{\Delta}(x)-f_{\Delta}(y)\rangle+q_{0}\Delta|g_{\Delta}(x)-g_{\Delta}(y)|^{2}
+Δ2|fΔ(x)fΔ(y)|2+q0Δ2|ggΔ(x)ggΔ(y)|2\displaystyle\quad+\Delta^{2}|f_{\Delta}(x)-f_{\Delta}(y)|^{2}+q_{0}\Delta^{2}|g^{\prime}\cdot g_{\Delta}(x)-g^{\prime}\cdot g_{\Delta}(y)|^{2}
(1+CΔ)|xy|2+CΔ22γ(α^β^)|xy|2\displaystyle\leq(1+C\Delta)|x-y|^{2}+C\Delta^{2-2\gamma(\hat{\alpha}\vee\hat{\beta})}|x-y|^{2}
(1+CΔ)|xy|2.\displaystyle\leq(1+C\Delta)|x-y|^{2}.

Thus, the proof is finished. \Box

The following lemma is a consequence of the polynomial growth bound (3.64)-(3.65).

Lemma 3.9

Let Assumptions 3.7, 2.2 and 2.3 hold. Then,

sup0tT𝔼[|𝕃f(X(t))|2]sup0tT𝔼[|𝕃g(X(t))|2]C(1+sup0tT𝔼[|X(t)|4α^+2]+sup0tT𝔼[|X(t)|4β^]),\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[|\mathbb{L}f(X(t))|^{2}\right]\vee\sup_{0\leq t\leq T}\mathbb{E}\left[|\mathbb{L}g(X(t))|^{2}\right]\leq C\Big{(}1+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{4\hat{\alpha}+2}\big{]}+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{-4\hat{\beta}}\big{]}\Big{)}, (3.72)
sup0tT𝔼[|f(X(t))g(X(t))|2]C(1+sup0tT𝔼[|X(t)|3α^+2]+sup0tT𝔼[|X(t)|3β^]),\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[|f^{\prime}(X(t))g(X(t))|^{2}\right]\leq C\Big{(}1+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{3\hat{\alpha}+2}\big{]}+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{-3\hat{\beta}}\big{]}\Big{)}, (3.73)

From (3.8), we have

𝔼[|tt+Δ(f(X(s))f(X(t)))𝑑s|2]C(1+sup0tT𝔼[|X(t)|4α^+2]+sup0tT𝔼[|X(t)|4β^])Δ3.\displaystyle\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}f(X(s))-f(X(t))\Big{)}ds\Big{|}^{2}\right]\leq C\Big{(}1+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{4\hat{\alpha}+2}\big{]}+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{-4\hat{\beta}}\big{]}\Big{)}\Delta^{3}. (3.74)

However, if we insert the conditional expectation with respect to the σ\sigma-field t\mathcal{F}_{t}, the order convergence indicated by (3.74) can be increased, see (3.75).

Lemma 3.10

Let Assumptions 3.7, 2.2 and 2.3 hold with

p04α^+2andp14β^.\displaystyle p_{0}\geq 4\hat{\alpha}+2\quad\textrm{and}\quad p_{1}\geq 4\hat{\beta}.

Then for any t[0,T]t\in[0,T],

𝔼[|𝔼t[tt+Δ(f(X(s))f(X(t)))𝑑s]|2]=𝔼[|tt+Δ(ts𝕃f(X(u))𝑑u)𝑑s|2]CΔ4,\displaystyle\mathbb{E}\left[\Big{|}\mathbb{E}_{t}\Big{[}\int_{t}^{t+\Delta}\Big{(}f(X(s))-f(X(t))\Big{)}ds\Big{]}\Big{|}^{2}\right]=\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}\int_{t}^{s}\mathbb{L}f(X(u))du\Big{)}ds\Big{|}^{2}\right]\leq C\Delta^{4}, (3.75)
𝔼[|tt+Δ(tsf(X(u))g(X(u))𝑑B(u))𝑑s|2]CΔ3,\displaystyle\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}\int_{t}^{s}f^{\prime}(X(u))g(X(u))dB(u)\Big{)}ds\Big{|}^{2}\right]\leq C\Delta^{3}, (3.76)
𝔼[|tt+Δ(g(X(s))g(X(t)))𝑑B(s)g(X(t))g(X(t))It,t+Δ|2]CΔ3.\displaystyle\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}g(X(s))-g(X(t))\Big{)}dB(s)-g^{\prime}(X(t))\cdot g(X(t))I_{t,t+\Delta}\Big{|}^{2}\right]\leq C\Delta^{3}. (3.77)

Proof. By the Itô formula, we have

f(X(s))f(X(t))=ts𝕃f(X(u))𝑑u+tsf(X(u))g(X(u))𝑑B(u),0ts.\displaystyle f(X(s))-f(X(t))=\int_{t}^{s}\mathbb{L}f(X(u))du+\int_{t}^{s}f^{\prime}(X(u))g(X(u))dB(u),\quad 0\leq t\leq s. (3.78)

Thus,

𝔼t[tt+Δ(f(X(s))f(X(t)))𝑑s]=tt+Δ(ts𝕃f(X(u))𝑑u)𝑑s\displaystyle\mathbb{E}_{t}\Big{[}\int_{t}^{t+\Delta}\Big{(}f(X(s))-f(X(t))\Big{)}ds\Big{]}=\int_{t}^{t+\Delta}\Big{(}\int_{t}^{s}\mathbb{L}f(X(u))du\Big{)}ds (3.79)

By (3.72), we have

𝔼[|tt+Δ(ts𝕃f(X(u))𝑑u)𝑑s|2]Δ2tt+Δ(ts𝔼[|𝕃f(X(u))|2]𝑑u)𝑑sCΔ4,\displaystyle\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}\int_{t}^{s}\mathbb{L}f(X(u))du\Big{)}ds\Big{|}^{2}\right]\leq\Delta^{2}\int_{t}^{t+\Delta}\Big{(}\int_{t}^{s}\mathbb{E}[|\mathbb{L}f(X(u))|^{2}]du\Big{)}ds\leq C\Delta^{4},

which implies that (3.75) holds.

Moreover, by the Itô isometry and (3.73), we also have

𝔼[|tt+Δ(tsf(X(u))g(X(u))𝑑B(u))𝑑s|2]\displaystyle\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}\int_{t}^{s}f^{\prime}(X(u))g(X(u))dB(u)\Big{)}ds\Big{|}^{2}\right]
Δtt+Δ(𝔼[|tsf(X(u))g(X(u))𝑑B(u)|2])𝑑s\displaystyle\leq\Delta\int_{t}^{t+\Delta}\Big{(}\mathbb{E}\left[\Big{|}\int_{t}^{s}f^{\prime}(X(u))g(X(u))dB(u)\Big{|}^{2}\right]\Big{)}ds
=Δtt+Δts𝔼[|f(X(u))g(X(u))|2]𝑑u𝑑sCΔ3,\displaystyle=\Delta\int_{t}^{t+\Delta}\int_{t}^{s}\mathbb{E}\left[|f^{\prime}(X(u))g(X(u))|^{2}\right]duds\leq C\Delta^{3},

which is the desired assertion (3.76).

It remains to show (3.77). Again, by the Itô isometry, we have

𝔼[|tt+Δ(g(X(s))g(X(t)))𝑑B(s)gg(X(t))It,t+Δ|2]\displaystyle\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}g(X(s))-g(X(t))\Big{)}dB(s)-g^{\prime}\cdot g(X(t))I_{t,t+\Delta}\Big{|}^{2}\right]
=𝔼[|tt+Δ(g(X(s))g(X(t))gg(X(t))[B(s)B(t)])𝑑B(s)|2]\displaystyle=\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}\Big{(}g(X(s))-g(X(t))-g^{\prime}\cdot g(X(t))[B(s)-B(t)]\Big{)}dB(s)\Big{|}^{2}\right]
=tt+ΔΓ(s)𝑑s,\displaystyle=\int_{t}^{t+\Delta}\Gamma(s)ds, (3.80)

where

Γ(s):=𝔼[|g(X(s))g(X(t))gg(X(t))[B(s)B(t)]|2].\Gamma(s):=\mathbb{E}\left[\Big{|}g(X(s))-g(X(t))-g^{\prime}\cdot g(X(t))[B(s)-B(t)]\Big{|}^{2}\right].

Thus, the assertion (3.77) is proved if there is a constant CC independent of s,ts,t and Δ\Delta such that

Γ(s)CΔ2,s[t,t+Δ].\displaystyle\Gamma(s)\leq C\Delta^{2},\quad\forall s\in[t,t+\Delta]. (3.81)

Again, by the Itô formula, we get

g(X(s))g(X(t))=ts𝕃g(X(u))𝑑u+tsgg(X(u))𝑑B(u),0ts.\displaystyle g(X(s))-g(X(t))=\int_{t}^{s}\mathbb{L}g(X(u))du+\int_{t}^{s}g^{\prime}\cdot g(X(u))dB(u),\quad 0\leq t\leq s.

Thus,

Γ(s)\displaystyle\Gamma(s) =𝔼[|ts𝕃g(X(u))𝑑u+ts(gg(X(u))gg(X(t)))𝑑B(s)|2]\displaystyle=\mathbb{E}\left[\Big{|}\int_{t}^{s}\mathbb{L}g(X(u))du+\int_{t}^{s}\Big{(}g^{\prime}\cdot g(X(u))-g^{\prime}\cdot g(X(t))\Big{)}dB(s)\Big{|}^{2}\right]
2𝔼[|ts𝕃g(X(u))𝑑u|2]:=Γ1(s)+2𝔼[|ts(gg(X(u))gg(X(t)))𝑑B(s)|2]:=Γ2(s).\displaystyle\leq\underbrace{2\mathbb{E}\left[\Big{|}\int_{t}^{s}\mathbb{L}g(X(u))du\Big{|}^{2}\right]}_{:=\Gamma_{1}(s)}+\underbrace{2\mathbb{E}\left[\Big{|}\int_{t}^{s}\Big{(}g^{\prime}\cdot g(X(u))-g^{\prime}\cdot g(X(t))\Big{)}dB(s)\Big{|}^{2}\right]}_{:=\Gamma_{2}(s)}. (3.82)

From (3.72), we have

Γ1(s)2Δts𝔼[|𝕃g(X(u))|2]𝑑sCΔ2.\displaystyle\Gamma_{1}(s)\leq 2\Delta\int_{t}^{s}\mathbb{E}\left[|\mathbb{L}g(X(u))|^{2}\right]ds\leq C\Delta^{2}. (3.83)

By the Itô isometry, we get

Γ2(s)=2ts𝔼[|(gg(X(u))gg(X(t)))|2]𝑑s.\displaystyle\Gamma_{2}(s)=2\int_{t}^{s}\mathbb{E}\left[\Big{|}\Big{(}g^{\prime}\cdot g(X(u))-g^{\prime}\cdot g(X(t))\Big{)}\Big{|}^{2}\right]ds. (3.84)

In the similar fashion as (3.7) is obtained, we also can show that

𝔼[|(gg(X(u))gg(X(t)))|2]\displaystyle\mathbb{E}\left[\Big{|}\Big{(}g^{\prime}\cdot g(X(u))-g^{\prime}\cdot g(X(t))\Big{)}\Big{|}^{2}\right]
C(1+sup0tT𝔼[|X(t)|4α^+2]+sup0tT𝔼[|X(t)|4β^])Δ,tust+Δ.\displaystyle\leq C\Big{(}1+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{4\hat{\alpha}+2}\big{]}+\sup_{0\leq t\leq T}\mathbb{E}\big{[}|X(t)|^{-4\hat{\beta}}\big{]}\Big{)}\Delta,\quad t\leq u\leq s\leq t+\Delta. (3.85)

Thus, inserting (3.84) and (3.83) into (3.2), we conclude that Γ(s)CΔ2,s[t,t+Δ]\Gamma(s)\leq C\Delta^{2},\quad\forall s\in[t,t+\Delta], which completes the proof of (3.77). \Box

Lemma 3.11

Let Assumptions 3.7, 2.2 and 2.3 hold with

p04(α^β^)+2α^+2andp14(α^β^)+2β^2.\displaystyle p_{0}\geq 4(\hat{\alpha}\vee\hat{\beta})+2\hat{\alpha}+2\quad\textrm{and}\quad p_{1}\geq 4(\hat{\alpha}\vee\hat{\beta})+2\hat{\beta}-2. (3.86)

Then

𝔼[|𝔼tk[M^k(d)]|2]\displaystyle\mathbb{E}\left[\big{|}\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\right] CΔ4,k=0,1,,\displaystyle\leq C\Delta^{4},\quad k=0,1,\cdots, (3.87)
𝔼[|M^k(w)|2]\displaystyle\mathbb{E}\left[\big{|}\hat{M}^{(w)}_{k}\big{|}^{2}\right] CΔ3,k=0,1,,\displaystyle\leq C\Delta^{3},\quad k=0,1,\cdots, (3.88)

where

M^k(d):=tktk+1(f(X(s))fΔ(X(tk)))𝑑s,\displaystyle\hat{M}^{(d)}_{k}:=\int_{t_{k}}^{t_{k+1}}\Big{(}f(X(s))-f_{\Delta}(X(t_{k}))\Big{)}ds,
M^k(w):=tktk+1(g(X(s))gΔ(X(tk)))𝑑B(s)ggΔ(X(tk))Itk,tk+1,\displaystyle\hat{M}^{(w)}_{k}:=\int_{t_{k}}^{t_{k+1}}\Big{(}g(X(s))-g_{\Delta}(X(t_{k}))\Big{)}dB(s)-g^{\prime}\cdot g_{\Delta}(X(t_{k}))I_{t_{k},t_{k+1}},

with R(Δ)=L1Δ12(α^β^)R(\Delta)=L_{1}\Delta^{-\frac{1}{2(\hat{\alpha}\vee\hat{\beta})}} and Itk,tk+1=12(|ΔBk|2Δ)I_{t_{k},t_{k+1}}=\frac{1}{2}(|\Delta B_{k}|^{2}-\Delta).

Proof. Let condition (3.86) hold and set γ=12(α^β^)\gamma=\frac{1}{2(\hat{\alpha}\vee\hat{\beta})}. Then

γ(p02α^2)2andγ(p12β^+2)2.\displaystyle\gamma(p_{0}-2\hat{\alpha}-2)\geq 2\quad\textrm{and}\quad\gamma(p_{1}-2\hat{\beta}+2)\geq 2. (3.89)

Thus, we conclude from Lemma 2.6 that

sup0tT𝔼[|f(X(t))fΔ(X(t)|p]sup0tT𝔼[|g(X(t))gΔ(X(t)|p]\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[|f(X(t))-f_{\Delta}(X(t)|^{p}\right]\vee\sup_{0\leq t\leq T}\mathbb{E}\left[|g(X(t))-g_{\Delta}(X(t)|^{p}\right]
C(Δγ(p02α^2)+Δγ(p12β^+2))CΔ2.\displaystyle\leq C\Big{(}\Delta^{\gamma(p_{0}-2\hat{\alpha}-2)}+\Delta^{\gamma(p_{1}-2\hat{\beta}+2)}\Big{)}\leq C\Delta^{2}. (3.90)

Moreover, we have the following decomposition

M^k(d)=tktk+1(f(X(s))f(X(tk)))𝑑s+(f(X(tk))fΔ(X(tk)))Δ.\displaystyle\hat{M}^{(d)}_{k}=\int_{t_{k}}^{t_{k+1}}\Big{(}f(X(s))-f(X(t_{k}))\Big{)}ds+\Big{(}f(X(t_{k}))-f_{\Delta}(X(t_{k}))\Big{)}\Delta.

Therefore,

𝔼[|𝔼tk[M^k(d)]|2]\displaystyle\mathbb{E}\left[\big{|}\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\right]
2𝔼[|𝔼tk[tktk+Δ(f(X(s))f(X(tk)))𝑑s]|2]+2Δ2𝔼[|f(X(tk))fΔ(X(tk))|2]\displaystyle\leq 2\mathbb{E}\left[\Big{|}\mathbb{E}_{t_{k}}\Big{[}\int_{t_{k}}^{t_{k}+\Delta}\Big{(}f(X(s))-f(X(t_{k}))\Big{)}ds\Big{]}\Big{|}^{2}\right]+2\Delta^{2}\mathbb{E}\left[\Big{|}f(X(t_{k}))-f_{\Delta}(X(t_{k}))\Big{|}^{2}\right]
CΔ4+CΔ4=CΔ4,\displaystyle\leq C\Delta^{4}+C\Delta^{4}=C\Delta^{4},

where (3.75) and (3.2) have been used. In addition, Itô isometry implies that

𝔼[|M^k(w)|2]\displaystyle\mathbb{E}\left[\big{|}\hat{M}^{(w)}_{k}\big{|}^{2}\right] =𝔼[|tktk+1(g(X(s))gΔ(X(tk))ggΔ(X(tk))[B(s)B(tk)])𝑑B(s)|2]\displaystyle=\mathbb{E}\left[\Big{|}\int_{t_{k}}^{t_{k+1}}\Big{(}g(X(s))-g_{\Delta}(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(X(t_{k}))[B(s)-B(t_{k})]\Big{)}dB(s)\Big{|}^{2}\right]
=tktk+1Γ^(s)𝑑s,\displaystyle=\int_{t_{k}}^{t_{k+1}}\hat{\Gamma}(s)ds, (3.91)

where

Γ^(s):=𝔼[|g(X(s))gΔ(X(tk))ggΔ(X(tk))[B(s)B(tk)]|2].\displaystyle\hat{\Gamma}(s):=\mathbb{E}\left[\big{|}g(X(s))-g_{\Delta}(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(X(t_{k}))[B(s)-B(t_{k})]\big{|}^{2}\right].

By (3.81) and (2.17), we have

Γ^(s)\displaystyle\hat{\Gamma}(s) 3𝔼[|g(X(s))g(X(tk))gg(X(tk))[B(s)B(tk)]|2]\displaystyle\leq 3\mathbb{E}\left[\big{|}g(X(s))-g(X(t_{k}))-g^{\prime}\cdot g(X(t_{k}))[B(s)-B(t_{k})]\big{|}^{2}\right]
+3𝔼[|g(X(tk))gΔ(X(tk))|2]+3𝔼[|gg(X(tk))ggΔ(X(tk))|2|B(s)B(tk)|2]\displaystyle\quad+3\mathbb{E}\left[\big{|}g(X(t_{k}))-g_{\Delta}(X(t_{k}))\big{|}^{2}\right]+3\mathbb{E}\left[\big{|}g^{\prime}\cdot g(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(X(t_{k}))\big{|}^{2}\big{|}B(s)-B(t_{k})\big{|}^{2}\right]
CΔ2+C(Δγ(p02α^2)+Δγ(p12β^+2))+CΔ𝔼[|gg(X(tk))ggΔ(X(tk))|2],\displaystyle\leq C\Delta^{2}+C\Big{(}\Delta^{\gamma(p_{0}-2\hat{\alpha}-2)}+\Delta^{\gamma(p_{1}-2\hat{\beta}+2)}\Big{)}+C\Delta\mathbb{E}\left[\big{|}g^{\prime}\cdot g(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(X(t_{k}))\big{|}^{2}\right],
CΔ2+CΔ𝔼[|gg(X(tk))ggΔ(X(tk))|2],\displaystyle\leq C\Delta^{2}+C\Delta\mathbb{E}\left[\big{|}g^{\prime}\cdot g(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(X(t_{k}))\big{|}^{2}\right],

where the last step follows from (3.89). In a similar fashion as (2.17) was proved, we also can show that

sup0tT𝔼[|gg(X(t))ggΔ(X(t))|2]C(Δγ(p02α^2)+Δγ(p12β^+2))CΔ2,\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[\big{|}g^{\prime}\cdot g(X(t))-g^{\prime}\cdot g_{\Delta}(X(t))\big{|}^{2}\right]\leq C\Big{(}\Delta^{\gamma(p_{0}-2\hat{\alpha}-2)}+\Delta^{\gamma(p_{1}-2\hat{\beta}+2)}\Big{)}\leq C\Delta^{2},

where (3.89) has been used. Thus,

Γ^(s)CΔ2.\hat{\Gamma}(s)\leq C\Delta^{2}.

Inserting this into (3.2) completes the proof of (3.88). \Box

The following theorem shows that the TMil achieves the optimal mean-square convergence order 11.

Theorem 3.12 (Convergence order of TMil)

Let Assumptions 3.7, 2.2 and 2.3 hold with

p04(α^β^)+2α^+2andp14(α^β^)+2β^.\displaystyle p_{0}\geq 4(\hat{\alpha}\vee\hat{\beta})+2\hat{\alpha}+2\quad\textrm{and}\quad p_{1}\geq 4(\hat{\alpha}\vee\hat{\beta})+2\hat{\beta}. (3.92)

Then, the TMil scheme (3.69) by setting

R(Δ)=L1Δ12(α^β^)\displaystyle R(\Delta)=L_{1}\Delta^{-\frac{1}{2(\hat{\alpha}\vee\hat{\beta})}} (3.93)

has the property that

max0kT/Δ𝔼[|X(tk)X^Δ(tk)|2]CΔ2,T>0,\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\hat{X}_{\Delta}(t_{k})|^{2}]\leq C\Delta^{2},\quad T>0, (3.94)
max0kT/Δ𝔼[|X(tk)πΔ(X^Δ(tk))|2]CΔ2,T>0,\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{2}]\leq C\Delta^{2},\quad T>0, (3.95)

where CC is a positive constant independent of Δ\Delta.

Proof. We observe from (2.2) that

X(tk+1)=X(tk)+fΔ(X(tk))Δ+gΔ(X(tk))ΔBk+ggΔ(X(tk))Itk,tk+1+M^k(d)+M^k(w),\displaystyle X(t_{k+1})=X(t_{k})+f_{\Delta}(X(t_{k}))\Delta+g_{\Delta}(X(t_{k}))\Delta B_{k}+g^{\prime}\cdot g_{\Delta}(X(t_{k}))I_{t_{k},t_{k+1}}+\hat{M}^{(d)}_{k}+\hat{M}^{(w)}_{k}, (3.96)

where M^k(d)\hat{M}^{(d)}_{k} and M^k(w)\hat{M}^{(w)}_{k} are defined in Lemma 3.11. Thus, we conclude from (3.70) and (3.96) that

e^k+1Δ:=X(tk+1)X^Δ(tk+1)\displaystyle\hat{e}^{\Delta}_{k+1}:=X(t_{k+1})-\hat{X}_{\Delta}(t_{k+1}) (3.97)
=[X(tk)X^Δ(tk)]+[fΔ(X(tk))fΔ(X^Δ(tk))]Δ+[gΔ(X(tk))gΔ(X^Δ(tk))]ΔBk\displaystyle=[X(t_{k})-\hat{X}_{\Delta}(t_{k})]+[f_{\Delta}(X(t_{k}))-f_{\Delta}(\hat{X}_{\Delta}(t_{k}))]\Delta+[g_{\Delta}(X(t_{k}))-g_{\Delta}(\hat{X}_{\Delta}(t_{k}))]\Delta B_{k}
+[ggΔ(X(tk))ggΔ(X^Δ(tk))]Itk,tk+1+M^k(d)+M^k(w).\displaystyle\quad+\big{[}g^{\prime}\cdot g_{\Delta}(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(\hat{X}_{\Delta}(t_{k}))\big{]}I_{t_{k},t_{k+1}}+\hat{M}^{(d)}_{k}+\hat{M}^{(w)}_{k}. (3.98)

By the orthogonality of the conditional expectation it holds

𝔼[|ek+1Δ|2]=𝔼[|𝔼tk[ek+1Δ]|2]+𝔼[|ek+1Δ𝔼tk[ek+1Δ]|2]\displaystyle\mathbb{E}\left[\big{|}e^{\Delta}_{k+1}\big{|}^{2}\right]=\mathbb{E}\left[\big{|}\mathbb{E}_{t_{k}}[e^{\Delta}_{k+1}]\big{|}^{2}\right]+\mathbb{E}\left[\big{|}e^{\Delta}_{k+1}-\mathbb{E}_{t_{k}}[e^{\Delta}_{k+1}]\big{|}^{2}\right]

Thus, by the elementary inequality

(a+b)2(1+ε)a2+(1+1/ε)b2,a,b,ε>0,\displaystyle(a+b)^{2}\leq(1+\varepsilon)a^{2}+(1+1/\varepsilon)b^{2},\quad a,b,\varepsilon>0, (3.99)

and

𝔼[|Itk,tk+1|2]=12𝔼[|ΔBkΔ|2]=Δ22,\mathbb{E}\big{[}|I_{t_{k},t_{k+1}}|^{2}\big{]}=\frac{1}{2}\mathbb{E}\Big{[}|\Delta B_{k}-\Delta|^{2}\Big{]}=\frac{\Delta^{2}}{2},

as well as Lemma 3.8, we have

𝔼[|X(tk+1)X^Δ(tk+1)|2]\displaystyle\mathbb{E}\left[\big{|}X(t_{k+1})-\hat{X}_{\Delta}(t_{k+1})\big{|}^{2}\right] (3.100)
=𝔼[|[X(tk)X^Δ(tk)]+[fΔ(X(tk))fΔ(X^Δ(tk))]Δ+𝔼tk[M^k(d)]|2]+\displaystyle=\mathbb{E}\left[\big{|}[X(t_{k})-\hat{X}_{\Delta}(t_{k})]+[f_{\Delta}(X(t_{k}))-f_{\Delta}(\hat{X}_{\Delta}(t_{k}))]\Delta+\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\right]+
+𝔼[|[gΔ(X(tk))gΔ(X^Δ(tk))]ΔBk+[ggΔ(X(tk))ggΔ(X^Δ(tk))]Itk,tk+1\displaystyle\quad+\mathbb{E}\Big{[}\big{|}[g_{\Delta}(X(t_{k}))-g_{\Delta}(\hat{X}_{\Delta}(t_{k}))]\Delta B_{k}+\big{[}g^{\prime}\cdot g_{\Delta}(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(\hat{X}_{\Delta}(t_{k}))\big{]}I_{t_{k},t_{k+1}}
+M^k(w)+(M^k(d)𝔼tk[M^k(d)])|2]\displaystyle\quad+\hat{M}^{(w)}_{k}+\big{(}\hat{M}^{(d)}_{k}-\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{)}\big{|}^{2}\Big{]}
(1+Δ)𝔼[|[X(tk)X^Δ(tk)]+Δ[fΔ(X(tk))fΔ(X^Δ(tk))]|2]\displaystyle\leq(1+\Delta)\mathbb{E}\left[\big{|}[X(t_{k})-\hat{X}_{\Delta}(t_{k})]+\Delta[f_{\Delta}(X(t_{k}))-f_{\Delta}(\hat{X}_{\Delta}(t_{k}))]\big{|}^{2}\right]
+q0Δ𝔼[|gΔ(X(tk))gΔ(X^Δ(tk))|2]+q0Δ2𝔼[|ggΔ(X(tk))ggΔ(X^Δ(tk))|2]\displaystyle\quad+q_{0}\Delta\mathbb{E}\big{[}|g_{\Delta}(X(t_{k}))-g_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{2}\big{]}+q_{0}\Delta^{2}\mathbb{E}\big{[}|g^{\prime}\cdot g_{\Delta}(X(t_{k}))-g^{\prime}\cdot g_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{2}\big{]}
+(1+1Δ)𝔼[|𝔼tk[M^k(d)]|2]+C𝔼[|M^k(w)|2]+C𝔼[|M^k(d)𝔼tk[M^k(d)]|2]\displaystyle\quad+\left(1+\frac{1}{\Delta}\right)\mathbb{E}\left[\big{|}\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\right]+C\mathbb{E}\big{[}\big{|}\hat{M}^{(w)}_{k}\big{|}^{2}\big{]}+C\mathbb{E}\big{[}\big{|}\hat{M}^{(d)}_{k}-\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\big{]}
(1+CΔ)𝔼[|X(tk)X^Δ(tk)|2]+Ξk,\displaystyle\leq(1+C\Delta)\mathbb{E}\left[\big{|}X(t_{k})-\hat{X}_{\Delta}(t_{k})\big{|}^{2}\right]+\Xi_{k}, (3.101)

where

Ξk:=(1+1Δ)𝔼[|𝔼tk[M^k(d)]|2]+C𝔼[|M^k(w)|2]+C𝔼[|M^k(d)𝔼tk[M^k(d)]|2].\displaystyle\Xi_{k}:=\left(1+\frac{1}{\Delta}\right)\mathbb{E}\left[\big{|}\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\right]+C\mathbb{E}\big{[}\big{|}\hat{M}^{(w)}_{k}\big{|}^{2}\big{]}+C\mathbb{E}\big{[}\big{|}\hat{M}^{(d)}_{k}-\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\big{]}.

According to (3.76), we have

𝔼[|M^k(d)𝔼tk[M^k(d)]|2]=𝔼[|tktk+Δ(tksf(X(u))g(X(u))𝑑B(u))𝑑s|2]CΔ3.\displaystyle\mathbb{E}\big{[}\big{|}\hat{M}^{(d)}_{k}-\mathbb{E}_{t_{k}}[\hat{M}^{(d)}_{k}]\big{|}^{2}\big{]}=\mathbb{E}\left[\Big{|}\int_{t_{k}}^{t_{k}+\Delta}\Big{(}\int_{t_{k}}^{s}f^{\prime}(X(u))g(X(u))dB(u)\Big{)}ds\Big{|}^{2}\right]\leq C\Delta^{3}.

Combining this and Lemma 3.11, we get

ΞkCΔ3.\Xi_{k}\leq C\Delta^{3}.

Inserting this into (3.100), we have

𝔼[|X(tk+1)X^Δ(tk+1)|2](1+CΔ)𝔼[|X(tk)X^Δ(tk)|2]+CΔ3,k=0,1,.\displaystyle\mathbb{E}\left[\big{|}X(t_{k+1})-\hat{X}_{\Delta}(t_{k+1})\big{|}^{2}\right]\leq(1+C\Delta)\mathbb{E}\left[\big{|}X(t_{k})-\hat{X}_{\Delta}(t_{k})\big{|}^{2}\right]+C\Delta^{3},\quad k=0,1,\cdots.

Using the discrete Gronwall inequality yields the desired assertion (3.94). It remains to prove (3.95). From Lemma 2.6, we get immediately that

max0kT/Δ𝔼[|X(tk)πΔ(X(tk))|2]C(Δγ(p02)+Δγ(p1+2))CΔ2.\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[\big{|}X(t_{k})-\pi_{\Delta}(X(t_{k}))\big{|}^{2}\right]\leq C\big{(}\Delta^{\gamma(p_{0}-2)}+\Delta^{\gamma(p_{1}+2)}\big{)}\leq C\Delta^{2}.

Combining this and Lemma 2.4 as well as the triangle inequality, we obtain the assertion (3.95). Thus, the proof is finished. \Box

If we apply Proposition 3.4 with p=2p=2, γ=12(α^β^)\gamma=\frac{1}{2(\hat{\alpha}\vee\hat{\beta})} and 𝔼[|ekΔ|p]CΔ2\mathbb{E}\left[|e^{\Delta}_{k}|^{p}\right]\leq C\Delta^{2}, then we have the following moment boundedness of the TMil solutions.

Proposition 3.13 (Moment boundedness of TMil)

Under the same assumption as Theorem 3.12, the TMil scheme (3.69) by setting (3.93) has the property that

  1. (1).

    For 2q4(α^β^)+1(p02α^1)(p12β^1)2\leq q\leq 4(\hat{\alpha}\vee\hat{\beta})+1\leq(p_{0}-2\hat{\alpha}-1)\wedge(p_{1}-2\hat{\beta}-1),

    max0kT/Δ𝔼[|πΔ(X^Δ(tk))|q]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{q}] <,T>0;\displaystyle<\infty,\quad T>0; (3.102)
  2. (2).

    For 1q4(α^β^)3(p02α^5)(p12β^6)1\leq q\leq 4(\hat{\alpha}\vee\hat{\beta})-3\leq(p_{0}-2\hat{\alpha}-5)\wedge(p_{1}-2\hat{\beta}-6),

    max0kT/Δ𝔼[|πΔ(X^Δ(tk))|q]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{-q}] <,T>0.\displaystyle<\infty,\quad T>0. (3.103)

4 Applications

We now apply our results to some SDEs in mathematical finance.

4.1 3/2 model

The 3/2 process XX is the solution to (1.1) and is strictly positive almost surely. Introduce the quantity

λ:=2+2c1σ2.\displaystyle\lambda:=2+\frac{2c_{1}}{\sigma^{2}}. (4.1)

Existence and uniqueness can be retrieved from the properties of the Feller diffusion, and

sup0tT𝔼|X(t)|p<,p<λ,\displaystyle\sup_{0\leq t\leq T}\mathbb{E}|X(t)|^{p}<\infty,\quad\forall p<\lambda, (4.2)

see [5, p. 1009]. Clearly, Assumption 2.1 is satisfied with exponents α=1\alpha=1 and β=0\beta=0. If condition

λ98q0+2,i.e., 4c1σ294q0\displaystyle\lambda\geq\frac{9}{8}q_{0}+2,\quad\textrm{i.e., }\quad\frac{4c_{1}}{\sigma^{2}}\geq\frac{9}{4}q_{0} (4.3)

holds, then

2f(x)+q0|g(x)|2=2c1c2(4c194q0σ2)x22c1c2,\displaystyle 2f^{\prime}(x)+q_{0}|g^{\prime}(x)|^{2}=2c_{1}c_{2}-\left(4c_{1}-\frac{9}{4}q_{0}\sigma^{2}\right)x^{2}\leq 2c_{1}c_{2},

which means that Assumption 2.2 holds. If λ>6\lambda>6, we choose p0[6,λ)p_{0}\in[6,\lambda), q0=2q_{0}=2 and fix γ=1p04\displaystyle\gamma=\frac{1}{p_{0}-4}, so that conditions 3.10 and 4.3 are satisfied. The following results reveal that the TEM for 3/2 model has L2L^{2}-order of 1/21/2 , 𝔼[|πΔ(XΔ(tk))|p]\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}\right] and 𝔼[|πΔ(XΔ(tk))|p]\mathbb{E}\left[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-p}\right] are finite from Proposition 3.5.

Corollary 4.1 (Convergence rate and moment boundedness of TEM for 3/2 model)

Let λ>6\lambda>6, i.e., σ2<c1/2\sigma^{2}<c_{1}/2. Then the truncated EM scheme (3.1) for model (1.1) by setting

R(Δ)=L1Δ1λ4R(\Delta)=L_{1}\Delta^{-\frac{1}{\lambda-4}}

has the property that

max0kT/Δ𝔼[|X(tk)πΔ(XΔ(tk))|2]CΔ.\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{2}]\leq C\Delta. (4.4)
  1. (1).

    For any p[2,λ3)p\in[2,\lambda-3),

    max0kT/Δ𝔼[|πΔ(XΔ(tk))|p]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}] <.\displaystyle<\infty. (4.5)
  2. (2).

    If λ>8\lambda>8, then for any p[1,λ7)p\in[1,\lambda-7),

    max0kT/Δ𝔼[|πΔ(XΔ(tk))|p]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{-p}] <.\displaystyle<\infty. (4.6)

Note that for model (1.1), Assumption 3.7 is satisfied with α^=1\hat{\alpha}=1 and β^=0\hat{\beta}=0. The following assertion follows directly from an application of Theorem 3.12 and Proposition 3.13.

Corollary 4.2 (Convergence rate of TMil for 3/2 model)

Let λ>8\lambda>8, i.e., σ2<c1/3\sigma^{2}<c_{1}/3. Then the truncated Milstein scheme (3.69) for model (1.1) by setting

R(Δ)=L1Δ12R(\Delta)=L_{1}\Delta^{-\frac{1}{2}}

has the property that

max0kT/Δ𝔼[|X(tk)X^Δ(tk)|2]CΔ2,\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\hat{X}_{\Delta}(t_{k})|^{2}]\leq C\Delta^{2}, (4.7)
max0kT/Δ𝔼[|X(tk)πΔ(X^Δ(tk))|2]CΔ2.\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|X(t_{k})-\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{2}]\leq C\Delta^{2}. (4.8)

Moreover,

max0kT/Δ𝔼[|πΔ(X^Δ(tk))|λ3]max0kT/Δ𝔼[|πΔ(X^Δ(tk))|(λ7)]<.\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{\lambda-3}]\vee\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}[|\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{-{(\lambda-7)}}]<\infty.

4.2 Aït-Sahalia model

Let XX be the solution to Aït-Sahalia interest rate model (1.2). If κ+1>2θ\kappa+1>2\theta, then there exists a strong solution on (0,)(0,\infty); sup0tT𝔼|X(t)|p\sup_{0\leq t\leq T}\mathbb{E}|X(t)|^{p} and sup0tT𝔼|X(t)|p\sup_{0\leq t\leq T}\mathbb{E}|X(t)|^{-p} are finite for any p>0p>0, see [27, Lemma 2.1]. In other words, Assumption 2.3 holds. Moreover, Assumptions 2.1 and 2.2 also hold for α=κ1\alpha=\kappa-1, β=2\beta=2 and q0>2q_{0}>2, see [24, p. 11]. The following results follows from Proposition 3.4.

Corollary 4.3 (Convergence rate of TEM for Aït-Sahalia model)

Let κ+1>2θ\kappa+1>2\theta with κ,θ>1\kappa,\theta>1. Then, for any p2p\geq 2, the TEM scheme (3.1) for model (1.2) by setting

R(Δ)=L1Δ1(2κ+2)8R(\Delta)=L_{1}\Delta^{-\frac{1}{(2\kappa+2)\vee 8}}

has the property that

max0kT/Δ𝔼[|X(tk)πΔ(XΔ(tk))|p]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}\right] CΔp/2,T>0,\displaystyle\leq C\Delta^{p/2},\quad T>0, (4.9)
max0kT/Δ𝔼[|πΔ(XΔ(tk))|p[(κ+1)4]+1]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\Big{[}|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p[(\kappa+1)\vee 4]+1}\Big{]} <,T>0,\displaystyle<\infty,\quad T>0, (4.10)
max0kT/Δ𝔼[1|πΔ(XΔ(tk))|p[(κ1)2]+1]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[\frac{1}{|\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p[(\kappa-1)\vee 2]+1}}\right] <,T>0.\displaystyle<\infty,\quad T>0. (4.11)

Clearly, Assumption 3.7 holds for α^=κ1\hat{\alpha}=\kappa-1 and β^=2\hat{\beta}=2. By Theorem 3.12 and Proposition 3.13, we have the following corollary.

Corollary 4.4 (Convergence rate of TMil for Aït-Sahalia model)

Under the same assumption as Corollary 4.3, the TMil scheme (3.69) for model (1.2) by setting

R(Δ)=L1Δ1(2κ2)4R(\Delta)=L_{1}\Delta^{-\frac{1}{(2\kappa-2)\vee 4}}

has the property that

max0kT/Δ𝔼[|X(tk)πΔ(X^Δ(tk))|2]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[|X(t_{k})-\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{2}\right] CΔ2,T>0.\displaystyle\leq C\Delta^{2},\quad T>0. (4.12)

For any p2p\geq 2,

max0kT/Δ𝔼[|πΔ(X^Δ(tk))|p]max0kT/Δ𝔼[|πΔ(X^Δ(tk))|p]<.\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[|\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{p}\right]\vee\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[|\pi_{\Delta}(\hat{X}_{\Delta}(t_{k}))|^{-p}\right]<\infty.
Remark 4.5

We remark that the authors in [24] have derived the strong convergence rate of order arbitrarily close to 12p\frac{1}{2p} of the TEM method for AIT model (1.2) for any p1p\geq 1 and ε(0,1/2)\varepsilon\in(0,1/2), i.e.,

max0kT/Δ𝔼[|X(tk)πΔ(XΔ(tk))|p]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[|X(t_{k})-\pi_{\Delta}(X_{\Delta}(t_{k}))|^{p}\right] CΔ1/2ε,\displaystyle\leq C\Delta^{1/2-\varepsilon}, (4.13)

see [24, Theorem 3.1]. In contrast, we achieve the optimal strong convergence rate of order 12\frac{1}{2}.

4.3 CIR model

The Cox-Ingersoll-Ross (CIR) process is given by the SDE

dX(t)=b1(b2X(t))dt+σX(t)dB(t),X(0)=X0>0,\displaystyle dX(t)=b_{1}(b_{2}-X(t))dt+\sigma\sqrt{X(t)}dB(t),\quad X(0)=X_{0}>0, (4.14)

where b1b_{1}, b2b_{2}, σ\sigma are strictly positive constant parameters. Under the Feller condition

ϖ:=2b1b2/σ2>1,\displaystyle\varpi:=2b_{1}b_{2}/\sigma^{2}>1, (4.15)

XX remains strictly positive almost surely. However Assumption 2.2 does not hold for CIR (4.14). Thus, our TEM can not apply to approximate CIR (4.14) directly. But, if we combine the Lamperti transform with our TEM, the strong LpL^{p}-convergence of order 1/21/2 can be derived for pp in a restricted parameter range.

Applying the Itô formula to the Lamperti transform Y=XY=\sqrt{X} gives a new SDE

dY(t)=f(Y(t))dtσ2dB(t),Y(0)=X0,\displaystyle dY(t)=f(Y(t))dt-\frac{\sigma}{2}dB(t),\quad Y(0)=\sqrt{X_{0}}, (4.16)

where

f(x)=a^x+b^x,a^=4b1b2σ28,b^:=b12.\displaystyle f(x)=\frac{\hat{a}}{x}+\hat{b}x,\quad\hat{a}=\frac{4b_{1}b_{2}-\sigma^{2}}{8},\quad\hat{b}:=-\frac{b_{1}}{2}.

From [33, p. 5], we have that

sup0tT𝔼[|X(t)|p]<,p>ϖ,\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[|X(t)|^{p}\right]<\infty,\quad\forall p>-\varpi,

and therefore,

sup0tT𝔼[|Y(t)|p]<,p<4b1b2σ2=2ϖ.\displaystyle\sup_{0\leq t\leq T}\mathbb{E}\left[|Y(t)|^{-p}\right]<\infty,\quad\forall p<\frac{4b_{1}b_{2}}{\sigma^{2}}=2\varpi.

Thus, for transformed SDE (4.16), Assumptions 2.1-2.3 hold with α=0\alpha=0, β=2\beta=2, p1<2ϖp_{1}<2\varpi, p0=+p_{0}=+\infty. Moreover, Assumption 3.7 also holds with α^=1\hat{\alpha}=1, β^=2\hat{\beta}=2. According to Proposition 3.4 and Theorem 3.12, we have the following results.

Corollary 4.6 (Convergence rate of Lamperti TEM/TMil for CIR model)

Let XX be the CIR process (4.14) and ϖ>5\varpi>5, where ϖ\varpi is the parameter defined by (4.15).

  • 1.

    Let YΔY_{\Delta} be the TEM solution (3.1) for (4.16) by setting R(Δ)=L1Δ18R(\Delta)=L_{1}\Delta^{-\frac{1}{8}}. Then for p[1,ϖ5]p\in\left[1,\frac{\varpi}{5}\right],

    max0kT/Δ𝔼[|X(tk)YΔ(tk)2|p]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[\big{|}X(t_{k})-Y_{\Delta}(t_{k})^{2}\big{|}^{p}\right] CΔp/2,T>0.\displaystyle\leq C\Delta^{p/2},\quad T>0. (4.17)
  • 2.

    Let Y^Δ\hat{Y}_{\Delta} be the TMil solution (3.69) for (4.16) by setting R(Δ)=L1Δ14R(\Delta)=L_{1}\Delta^{-\frac{1}{4}}. Then

    max0kT/Δ𝔼[|X(tk)Y^Δ(tk)2|]\displaystyle\max_{0\leq k\leq\lceil T/\Delta\rceil}\mathbb{E}\left[\big{|}X(t_{k})-\hat{Y}_{\Delta}(t_{k})^{2}\big{|}\right] CΔ,T>0.\displaystyle\leq C\Delta,\quad T>0. (4.18)

Proof. Let YY be the solution of (4.16). In order to approximate the original CIR process, we observe that

X(tk)YΔ(tk)2=Y(tk)2YΔ(tk)2=(Y(tk)+YΔ(tk))(Y(tk)YΔ(tk)).\displaystyle X(t_{k})-Y_{\Delta}(t_{k})^{2}=Y(t_{k})^{2}-Y_{\Delta}(t_{k})^{2}=\Big{(}Y(t_{k})+Y_{\Delta}(t_{k})\Big{)}\Big{(}Y(t_{k})-Y_{\Delta}(t_{k})\Big{)}.

Then the Hölder inequality gives

𝔼[|X(tk)YΔ(tk)2|p]=𝔼[|X(tk)+YΔ(tk)|p|X(tk)YΔ(tk)|p]\displaystyle\mathbb{E}\left[\big{|}X(t_{k})-Y_{\Delta}(t_{k})^{2}\big{|}^{p}\right]=\mathbb{E}\left[\big{|}X(t_{k})+Y_{\Delta}(t_{k})\big{|}^{p}\big{|}X(t_{k})-Y_{\Delta}(t_{k})\big{|}^{p}\right]
(𝔼[|X(tk)+YΔ(tk)|2p])1/2(𝔼[|X(tk)+YΔ(tk)|2p])1/2.\displaystyle\leq\left(\mathbb{E}\left[\big{|}X(t_{k})+Y_{\Delta}(t_{k})\big{|}^{2p}\right]\right)^{1/2}\left(\mathbb{E}\left[\big{|}X(t_{k})+Y_{\Delta}(t_{k})\big{|}^{2p}\right]\right)^{1/2}.

1pp110<ϖ51\leq p\leq\frac{p_{1}}{10}<\frac{\varpi}{5} implies that Condition (3.35) is satisfied. From (3.38), we obatian 𝔼[|YΔ(tk)|2p]<C\mathbb{E}\left[|Y_{\Delta}(t_{k})|^{2p}\right]<C. Similarly, 𝔼[|XΔ(tk)|2p]\mathbb{E}\left[|X_{\Delta}(t_{k})|^{2p}\right] is finite. This, combined with (3.37) lead to (4.17). Similarly, by Theorem 3.12 and Corollary 3.13, we can show that (4.18) also holds. \Box

Remark 4.7

Comparing Corollary 4.6 with Corollary 4.1 in [5], where the authors proved the strong LpL^{p}-convegence of order 1/p1/p of Lamperti projected EM for CIR, we observe that our Lamperti TEM has a significant improvement of the LpL^{p}-convergence order under slightly more stronger condtion on the parameters.

5 Numerical experiments

In this section we compare TEM (3.1) and TMil (3.70) to some numerical schemes that we outline below for the 3/2 and the AIT models.

  • 1.

    Euler-Maruyama scheme (EM) [34]:

    Yk+1=Yk+f(Yk)Δ+g(|Yk|)ΔBk,Y0=X0;Y_{k+1}=Y_{k}+f(Y_{k})\Delta+g(|Y_{k}|)\Delta B_{k},\quad Y_{0}=X_{0};
  • 2.

    Backward Euler-Maruyama scheme (BEM) [27]:

    Yk+1=Yk+f(Yk+1)Δ+g(|Yk|)ΔBk,Y0=X0;Y_{k+1}=Y_{k}+f(Y_{k+1})\Delta+g(|Y_{k}|)\Delta B_{k},\quad Y_{0}=X_{0};
  • 3.

    Logarithmic Truncated EM scheme (log TEM) [22]:

    Zk+1=Zk+F(π^(Zk+1))Δ+G(π^(Zk))ΔBk,Yk+1=eZk+1,Z0=log(X0),Y0=X0,Z_{k+1}=Z_{k}+F(\hat{\pi}(Z_{k+1}))\Delta+G(\hat{\pi}(Z_{k}))\Delta B_{k},\quad Y_{k+1}=\textrm{e}^{Z_{k+1}},\quad Z_{0}=\log(X_{0}),\quad Y_{0}=X_{0},

    where F(x)=exf(ex)0.5e2x|g(ex)|2F(x)=\textrm{e}^{-x}f(\textrm{e}^{x})-0.5\textrm{e}^{-2x}|g(\textrm{e}^{x})|^{2}, G(x)=exg(ex)G(x)=\textrm{e}^{-x}g(\textrm{e}^{x}), π^(x)=(R)xR\hat{\pi}(x)=(-R)\vee x\wedge R, R=C+logΔ1α(β+1)R=C+\frac{\log\Delta^{-1}}{\alpha\vee(\beta+1)};

  • 4.

    Semi-implicit Tamed EM scheme for AIT (STEM) [19]:

    Yk+1=Yk+a1Yk+11Δ+(a0+a1Yka2Ykκ1+ΔYkκ)Δ+bYkθ1+ΔYkκΔBk,Y0=X0,Y_{k+1}=Y_{k}+a_{-1}Y_{k+1}^{-1}\Delta+\left(-a_{0}+a_{1}Y_{k}-\frac{a_{2}Y_{k}^{\kappa}}{1+\sqrt{\Delta}Y_{k}^{\kappa}}\right)\Delta+\frac{bY_{k}^{\theta}}{1+\sqrt{\Delta}Y_{k}^{\kappa}}\Delta B_{k},\quad Y_{0}=X_{0},
  • 5.

    Semi-implicit Truncated EM scheme for AIT (STEM2) [18]:

    Yk+1=Yk+a1Yk+11Δ+(a0+a1Yka2Yˇkκ)Δ+bYˇkθΔBk,Y0=X0,Y_{k+1}=Y_{k}+a_{-1}Y_{k+1}^{-1}\Delta+\left(-a_{0}+a_{1}Y_{k}-a_{2}\check{Y}_{k}^{\kappa}\right)\Delta+b\check{Y}_{k}^{\theta}\Delta B_{k},\quad Y_{0}=X_{0},

    where Yˇk=(R)xR\check{Y}_{k}=(-R)\vee x\wedge R, R=Δ1/(2κ2)R=\Delta^{-1/(2\kappa-2)}.

Denote by Zj(T)Z^{\star}_{j}(T) the scheme \star approximation at time TT and by XjRef(T)X^{Ref}_{j}(T) the reference solution calculated by the corresponding scheme \star with small step size Δ=212\Delta=2^{-12}, using the same Brownian motion path (the jjth path). The root mean square error (RMSE) for the scheme \star is defined by

RMSE:=(1Mj=1M|XjRef(T)Zj(T)|2)1/2,\displaystyle\textrm{RMSE}^{\star}:=\left(\frac{1}{M}\sum_{j=1}^{M}|X^{Ref}_{j}(T)-Z^{\star}_{j}(T)|^{2}\right)^{1/2},

over MM sample paths. The strong error rates are computed by plotting RMSE against the step size on a log-log scale, and the strong rate of convergence is then retrieved using linear regression. Moreover, the event 𝐄Δ\mathbf{E}_{\Delta} is defined by

𝐄Δ:={min0kT/ΔXΔ(tk)0},\displaystyle\mathbf{E}_{\Delta}:=\left\{\min_{0\leq k\leq\lceil T/\Delta\rceil}X_{\Delta}(t_{k})\leq 0\right\}, (5.1)

where XΔX_{\Delta} is generated by the TEM scheme (3.2).

Example 5.1

Consider the following 3/2 model

dX(t)=(4X(t)4X(t)2)dt+σX(t)3/2dB(t),X(0)=2,\displaystyle dX(t)=(4X(t)-4X(t)^{2})dt+\sigma X(t)^{3/2}dB(t),\quad X(0)=2, (5.2)

where σ2<43\displaystyle\sigma^{2}<\frac{4}{3}. Now, we set T=2T=2, M=1000M=1000, R(Δ)=50Δ12R(\Delta)=50\Delta^{-\frac{1}{2}}. According to Corollaries 4.1 and 4.2, TEM solution YΔY_{\Delta} and TMil solution Y^Δ\hat{Y}_{\Delta} has the property that

𝔼|X(T)YΔ(T)|2CΔ,and𝔼|X(T)Y^Δ(T)|2CΔ2,\displaystyle\mathbb{E}|X(T)-Y_{\Delta}(T)|^{2}\leq C\Delta,\quad\textrm{and}\quad\mathbb{E}|X(T)-\hat{Y}_{\Delta}(T)|^{2}\leq C\Delta^{2},

respectively. RMSEs for different step sizes and rates of TEM and TMil for 3/2 model are displayed in Table 2 and Figure 1a. In the last line in Table 2, we observe the empirical rates 0.6138 and 1.0537 of TEM and TMil for σ=1/2\sigma=1/2, 0.6694 and 1.0613 of TEM and TMil for σ=1\sigma=1, which shows that the observed rates of convergence are slightly higher than the theoretical rates.

Moreover, the fourth and the seventh column in Table 2 lists respectively the probabilities of the TEM solutions escaping from +\mathbb{R}_{+} for different step sizes under a small noise intensity with σ=1/2\sigma=1/2, and a higher intensity with σ=1\sigma=1. We observe that as the step size goes to zero, this probability (𝐄Δ)\mathbb{P}(\mathbf{E}_{\Delta}) tends to zero. In other words, for a sufficiently small step size, say Δ=27\Delta=2^{-7}, the truncation πΔ\pi_{\Delta} from (3.1) does not execute for Example 5.1. In this context, our TEM and the EM coincide with a large probability.

Table 2: RMSEs and rates of TEM and TMil with different step sizes for 3/2 model
σ=1/2\sigma=1/2 σ=1\sigma=1
Δ\Delta TEM TMil (𝐄Δ)\mathbb{P}(\mathbf{E}_{\Delta}) TEM TMil (𝐄Δ)\mathbb{P}(\mathbf{E}_{\Delta})
232^{-3} 0.03%0.03\% 15.56%15.56\%
242^{-4} 0.00%0.00\% 2.90%2.90\%
252^{-5} 2.2170e-02 1.3748e-02 0.00%0.00\% 1.0827e-01 4.7061e-02 0.20%0.20\%
262^{-6} 1.3872e-02 6.3934e-03 0.00%0.00\% 5.4048e-02 2.0271e-02 0.01%0.01\%
272^{-7} 8.6830e-03 3.1602e-03 0.00%0.00\% 3.4534e-02 1.0442e-02 0.00%0.00\%
282^{-8} 5.9060e-03 1.4982e-03 0.00%0.00\% 2.3511e-02 5.0293e-03 0.00%0.00\%
292^{-9} 4.0491e-03 7.3669e-04 0.00%0.00\% 1.6132e-02 2.3878e-03 0.00%0.00\%
rate 0.6138 1.0537 0.6694 1.0613
00footnotetext: Note: This is an example of table footnote. This is an example of table footnote this is an example of table footnote this is an example of table footnote this is an example of table footnote.
Refer to caption
(a) RMSEs for 3/2 model with σ=1\sigma=1
Refer to caption
(b) RMSEs for AIT model
Figure 1: Convergence rates for 3/2 and AIT models
Table 3: Numerical results for different schemes for AIT
Δ\Delta log TEM STEM STEM2 BEM TEM TMil
252^{-5} 1.4201e-01 4.2682e-02 4.0054e-02 3.4322e-02 4.6424e-02 3.5164e-02
262^{-6} 1.0806e-01 3.1297e-02 2.5265e-02 2.6534e-02 2.7311e-02 1.5099e-02
272^{-7} 7.2431e-02 2.1658e-02 1.6477e-02 1.6354e-02 1.7300e-02 7.1460e-03
282^{-8} 5.3581e-02 1.4881e-02 1.1161e-02 1.1703e-02 1.1393e-02 3.2951e-03
292^{-9} 3.6888e-02 1.0145e-02 7.6605e-03 7.5196e-03 7.7554e-03 1.5746e-03
rate 0.4902 0.5218 0.5951 0.5566 0.6425 1.1158
CPU time 4.86s 4.19s 3.56s 2161.61s 3.16s 3.19s
00footnotetext: Note: This is an example of table footnote. This is an example of table footnote this is an example of table footnote this is an example of table footnote this is an example of table footnote.
Example 5.2

Consider the AIT model (1.2) with the following parameters

a1=1.5,a0=2,a1=1,a2=2,b=1,κ=4,θ=1.5,X0=1.\displaystyle a_{-1}=1.5,\quad a_{0}=2,\quad a_{1}=1,\quad a_{2}=2,\quad b=1,\quad\kappa=4,\quad\theta=1.5,\quad X_{0}=1.

Clearly, the above setting satisfies the condition κ+3>4θ\kappa+3>4\theta. Thus, all the methods listed in Table 1 can be applied to this AIT.

Sample trajectories from TEM, EM, log TEM, STEM, STEM2, and BEM for AIT are plotted in Figure 2, which shows that truncation occurs where the solution is close to zero and thus these methods except EM maintain positivity of the numerical approximations.

RMSEs and rates for different schemes are presented in Figure 1b and Table 3. In Figure 1b, we observe that log TEM has the largest error constant, whereas STEM, STEM2, BEM and TEM appear to have smaller error constants. Furthermore, the graphs of the log TEM, the STEM, the STEM2, and the BEM seem parallel to the reference line with slope equal to 0.50.5, while the TEM has a slightly bigger slope. Nevertheless, we observe in Figure 1b and Table 3 that TMil reaches the optimal convergence rate 11 and thus has a slight advantage over TEM when high accuracy is required. This behavior is also confirmed in the sixed line in Table 3. As expected the empirical rates of convergence are close to the theoretical ones.

Computational costs as measured by CPU time in seconds for difference methods are illustrated in the last line in Table 3, where we set M=104M=10^{4}, T=2T=2 and Δ=212\Delta=2^{-12}. One clearly observes that the explicit log TEM, STEM, STEM2, TEM and TMil greatly decreases the computational time compared to the implicit BEM. Since in the implementation of the BEM, we have to solve numerically a non-linear equation at each time step. This extra step brings questions about the computing performance of BEM.

Refer to caption
Figure 2: Sample trajectories computed with different schemes for AIT

6 Appendix

6.1 Proof of Lemma 2.4

Proof. Write R=R(Δ)R=R(\Delta) for simplicity.We first prove (2.10). For x,y(,R1]x,y\in(-\infty,R^{-1}] or [R1,R][R^{-1},R], or [R,)[R,\infty), obviously (2.10) holds. Without loss of generality, we assume that xyx\leq y. For R1xRyR^{-1}\leq x\leq R\leq y, πΔ(x)=x\pi_{\Delta}(x)=x, πΔ(y)=R\pi_{\Delta}(y)=R, then

|πΔ(x)πΔ(y)|=|xR|=Rxyx=|yx|.\displaystyle|\pi_{\Delta}(x)-\pi_{\Delta}(y)|=|x-R|=R-x\leq y-x=|y-x|. (6.1)

For xR1Ryx\leq R^{-1}\leq R\leq y, πΔ(x)=R1\pi_{\Delta}(x)=R^{-1}, πΔ(y)=R\pi_{\Delta}(y)=R,

|πΔ(x)πΔ(y)|=RR1|yx|.\displaystyle|\pi_{\Delta}(x)-\pi_{\Delta}(y)|=R-R^{-1}\leq|y-x|. (6.2)

For xR1yRx\leq R^{-1}\leq y\leq R, πΔ(x)=R1\pi_{\Delta}(x)=R^{-1}, πΔ(y)=y\pi_{\Delta}(y)=y,

|πΔ(x)πΔ(y)|=yR1|yx|.\displaystyle|\pi_{\Delta}(x)-\pi_{\Delta}(y)|=y-R^{-1}\leq|y-x|. (6.3)

From (6.1)-(6.3), we conclude that (2.10) holds.

Now, we begin to prove (2.11). For x,y[R,)x,y\in[R,\infty), or [R1,R][R^{-1},R], or (,R1](-\infty,R^{-1}], (2.11) holds obviously. Without loss of generality, we assume that xyx\leq y. For R1xRyR^{-1}\leq x\leq R\leq y, πΔ(x)=x\pi_{\Delta}(x)=x, πΔ(y)=R\pi_{\Delta}(y)=R, then by the Hölder inequality and Assumption 2.2, we have

2yx,fΔ(y)fΔ(x)+q0|gΔ(y)gΔ(x)|2\displaystyle 2\langle y-x,f_{\Delta}(y)-f_{\Delta}(x)\rangle+q_{0}|g_{\Delta}(y)-g_{\Delta}(x)|^{2} (6.4)
=2yx,f(R)f(x)+q0|g(R)g(x)|2\displaystyle=2\langle y-x,f(R)-f(x)\rangle+q_{0}|g(R)-g(x)|^{2}
=2(yx)xRf(u)𝑑u+q0|xRg(u)𝑑u|2\displaystyle=2(y-x)\int^{R}_{x}f^{\prime}(u)du+q_{0}\Big{|}\int^{R}_{x}g^{\prime}(u)du\Big{|}^{2}
(yR)xR2f(u)𝑑u+(Rx)xR2f(u)𝑑u+(Rx)xRq0|g(u)|2𝑑u\displaystyle\leq(y-R)\int^{R}_{x}2f^{\prime}(u)du+(R-x)\int^{R}_{x}2f^{\prime}(u)du+(R-x)\int^{R}_{x}q_{0}|g^{\prime}(u)|^{2}du
=(yR)xR2f(u)𝑑u+(Rx)xR2(f(u)+q0|g(u)|2)𝑑u\displaystyle=(y-R)\int^{R}_{x}2f^{\prime}(u)du+(R-x)\int^{R}_{x}2\Big{(}f^{\prime}(u)+q_{0}|g^{\prime}(u)|^{2}\Big{)}du
K(yR)(Rx)+K|Rx|2\displaystyle\leq K(y-R)(R-x)+K|R-x|^{2}
2K|yx|2.\displaystyle\leq 2K|y-x|^{2}.

Similarity, for xR1<Ryx\leq R^{-1}<R\leq y, πΔ(x)=R1\pi_{\Delta}(x)=R^{-1}, πΔ(y)=R\pi_{\Delta}(y)=R, we have

2yx,fΔ(y)fΔ(x)+q0|gΔ(y)gΔ(x)|2\displaystyle 2\langle y-x,f_{\Delta}(y)-f_{\Delta}(x)\rangle+q_{0}|g_{\Delta}(y)-g_{\Delta}(x)|^{2}
=2yx,f(R)f(R1)+q0|g(R)g(R1)|2\displaystyle=2\langle y-x,f(R)-f(R^{-1})\rangle+q_{0}|g(R)-g(R^{-1})|^{2}
=2(yx)R1Rf(u)𝑑u+q0|R1Rg(u)𝑑u|2\displaystyle=2(y-x)\int^{R}_{R^{-1}}f^{\prime}(u)du+q_{0}\Big{|}\int^{R}_{R^{-1}}g^{\prime}(u)du\Big{|}^{2}
(yR+R1x)R1R2f(u)𝑑u+(RR1)R1R2f(u)𝑑u+(RR1)R1Rq0|g(u)|2𝑑u\displaystyle\leq(y-R+R^{-1}-x)\int^{R}_{R^{-1}}2f^{\prime}(u)du+(R-R^{-1})\int^{R}_{R^{-1}}2f^{\prime}(u)du+(R-R^{-1})\int^{R}_{R^{-1}}q_{0}|g^{\prime}(u)|^{2}du
=(yR+R1x)R1R2f(u)𝑑u+(RR1)R1R2(f(u)+q0|g(u)|2)𝑑u\displaystyle=(y-R+R^{-1}-x)\int^{R}_{R^{-1}}2f^{\prime}(u)du+(R-R^{-1})\int^{R}_{R^{-1}}2\Big{(}f^{\prime}(u)+q_{0}|g^{\prime}(u)|^{2}\Big{)}du
K(yR+R1x)(Rx)+K|Rx|2\displaystyle\leq K(y-R+R^{-1}-x)(R-x)+K|R-x|^{2}
2K|yx|2.\displaystyle\leq 2K|y-x|^{2}. (6.5)

For xR1<yRx\leq R^{-1}<y\leq R, πΔ(x)=R1\pi_{\Delta}(x)=R^{-1}, πΔ(y)=y\pi_{\Delta}(y)=y, we have

2yx,fΔ(y)fΔ(x)+q0|gΔ(y)gΔ(x)|2\displaystyle 2\langle y-x,f_{\Delta}(y)-f_{\Delta}(x)\rangle+q_{0}|g_{\Delta}(y)-g_{\Delta}(x)|^{2}
=2yx,f(y)f(R1)+q0|g(y)g(R1)|2\displaystyle=2\langle y-x,f(y)-f(R^{-1})\rangle+q_{0}|g(y)-g(R^{-1})|^{2}
=2(yx)R1yf(u)𝑑u+q0|R1yg(u)𝑑u|2\displaystyle=2(y-x)\int^{y}_{R^{-1}}f^{\prime}(u)du+q_{0}\Big{|}\int^{y}_{R^{-1}}g^{\prime}(u)du\Big{|}^{2}
(R1x)R1y2f(u)𝑑u+(yR1)R1y2(f(u)+q0|g(u)|2)𝑑u\displaystyle\leq(R^{-1}-x)\int^{y}_{R^{-1}}2f^{\prime}(u)du+(y-R^{-1})\int^{y}_{R^{-1}}2\Big{(}f^{\prime}(u)+q_{0}|g^{\prime}(u)|^{2}\Big{)}du
K(R1x)(yR1)+K|yR1|2\displaystyle\leq K(R^{-1}-x)(y-R^{-1})+K|y-R^{-1}|^{2}
2K|yx|2.\displaystyle\leq 2K|y-x|^{2}.

Thus, the proof is finished. \Box

6.2 Proof of Lemma 2.5

Proof. For p[2,p01+αp1β]p\in\left[2,\frac{p_{0}}{1+\alpha}\wedge\frac{p_{1}}{\beta}\right] and t0t\geq 0, we conclude from Assumption 2.3 and (2.7) that

𝔼[|f(X(t))|p|g(X(t))|p]\displaystyle\mathbb{E}\left[|f(X(t))|^{p}\vee|g(X(t))|^{p}\right] C𝔼[(1+|X(t)|p(1+α)+|X(t)|pβ]\displaystyle\leq C\mathbb{E}\left[(1+|X(t)|^{p(1+\alpha)}+|X(t)|^{-p\beta}\right]
C(1+𝔼[|X(t)|p0]+𝔼[|X(t)|p1])C.\displaystyle\leq C(1+\mathbb{E}[|X(t)|^{p_{0}}]+\mathbb{E}[|X(t)|^{-p_{1}}])\leq C. (6.6)

Moreover,

X(t+Δ)X(t)=tt+Δf(X(s))𝑑s+tt+Δg(X(s))𝑑B(s).\displaystyle X(t+\Delta)-X(t)=\int_{t}^{t+\Delta}f(X(s))ds+\int_{t}^{t+\Delta}g(X(s))dB(s).

Hence, by the Itô isometry and (6.2) as well as (2.5), we have

𝔼[|X(t+Δ)X(t)|p]\displaystyle\mathbb{E}[|X(t+\Delta)-X(t)|^{p}] (6.7)
Cp(𝔼[|tt+Δf(X(s))𝑑s|p]+𝔼[|tt+Δg(X(s))𝑑B(s)|p])\displaystyle\leq C_{p}\left(\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}f(X(s))ds\Big{|}^{p}\right]+\mathbb{E}\left[\Big{|}\int_{t}^{t+\Delta}g(X(s))dB(s)\Big{|}^{p}\right]\right)
Cp(Δp1𝔼[tt+Δ|f(X(s))|p𝑑s]+Δp/21𝔼[tt+Δ|g(X(s))|p𝑑s])\displaystyle\leq C_{p}\left(\Delta^{p-1}\mathbb{E}\left[\int_{t}^{t+\Delta}\Big{|}f(X(s))\Big{|}^{p}ds\right]+\Delta^{p/2-1}\mathbb{E}\left[\int_{t}^{t+\Delta}\Big{|}g(X(s))\Big{|}^{p}ds\right]\right)
Cp(Δp1tt+Δsup0sT𝔼[|f(X(s))|p]ds+Δp/21tt+Δsup0sT𝔼[|g(X(s))|p]ds)\displaystyle\leq C_{p}\left(\Delta^{p-1}\int_{t}^{t+\Delta}\sup_{0\leq s\leq T}\mathbb{E}\left[\Big{|}f(X(s))\Big{|}^{p}\right]ds+\Delta^{p/2-1}\int_{t}^{t+\Delta}\sup_{0\leq s\leq T}\mathbb{E}\left[\Big{|}g(X(s))\Big{|}^{p}\right]ds\right)
CpΔp/2.\displaystyle\leq C_{p}\Delta^{p/2}.

For the case of p(0,2)p\in(0,2), the required results follows by the Lyapunov inequality. \Box

6.3 Proof of Lemma 2.6

Proof. Write R=R(Δ)R=R(\Delta) for simplicity. We conclude from (2.12) that

πΔ(x)+1πΔ(x)|x|+1|x|+1,x+.\displaystyle\pi_{\Delta}(x)+\frac{1}{\pi_{\Delta}(x)}\leq|x|+\frac{1}{|x|}+1,\quad\forall x\in\mathbb{R}_{+}. (6.8)

Combining this with (2.3), we observe that

|f(X(t))fΔ(X(t))|p|g(X(t))gΔ(X(t))|p\displaystyle|f(X(t))-f_{\Delta}(X(t))|^{p}\vee|g(X(t))-g_{\Delta}(X(t))|^{p} (6.9)
C(1+|X(t)|pα+|πΔ(X(t))|pα+|X(t)|pβ+|πΔ(X(t))|pβ)|X(t)πΔ(X(t))|p\displaystyle\leq C(1+|X(t)|^{p\alpha}+|\pi_{\Delta}(X(t))|^{p\alpha}+|X(t)|^{-p\beta}+|\pi_{\Delta}(X(t))|^{-p\beta})|X(t)-\pi_{\Delta}(X(t))|^{p}
C(1+|X(t)|pα+|X(t)|pβ)|X(t)πΔ(X(t))|p\displaystyle\leq C(1+|X(t)|^{p\alpha}+|X(t)|^{-p\beta})|X(t)-\pi_{\Delta}(X(t))|^{p}
C1Rp(1+1Rpα+1|X(t)|pβ)𝟏{X(t)<R1}+C|X(t)|p(1+|X(t)|pα+1Rpβ)𝟏{X(t)>R}\displaystyle\leq C\frac{1}{R^{p}}\Big{(}1+\frac{1}{R^{p\alpha}}+\frac{1}{|X(t)|^{p\beta}}\Big{)}\mathbf{1}_{\{X(t)<R^{-1}\}}+C|X(t)|^{p}\Big{(}1+|X(t)|^{p\alpha}+\frac{1}{R^{p\beta}}\Big{)}\mathbf{1}_{\{X(t)>R\}}
C1Rp(1+|X(t)|pβ)𝟏{X(t)<R1}=:J^1+C(1+|X(t)|pα+p)𝟏{X(t)>R}=:J^2.\displaystyle\leq\underbrace{C\frac{1}{R^{p}}\Big{(}1+|X(t)|^{-p\beta}\Big{)}\mathbf{1}_{\{X(t)<R^{-1}\}}}_{=:\hat{J}_{1}}+\underbrace{C\Big{(}1+|X(t)|^{p\alpha+p}\Big{)}\mathbf{1}_{\{X(t)>R\}}}_{=:\hat{J}_{2}}.

Consider first J^1\hat{J}_{1}. By the Hölder and Markov inequalities as well as Assumption 2.3, we have

𝔼[|X(t)|pβ𝟏{X(t)<R1}]\displaystyle\mathbb{E}\left[|X(t)|^{-p\beta}\mathbf{1}_{\{X(t)<R^{-1}\}}\right] C(𝔼[|X(t)|p1])pβ/p1({1/X(t)>R})1pβ/p1\displaystyle\leq C\Big{(}\mathbb{E}\left[|X(t)|^{-p_{1}}\right]\Big{)}^{p\beta/p_{1}}\Big{(}\mathbb{P}\{1/X(t)>R\}\Big{)}^{1-p\beta/p_{1}}
Cp1(𝔼[|X(t)|p1]Rp1)1pβ/p1Cp1R(Δ)p1pβ.\displaystyle\leq C_{p_{1}}\Big{(}\frac{\mathbb{E}\left[|X(t)|^{-p_{1}}\right]}{R^{p_{1}}}\Big{)}^{1-p\beta/p_{1}}\leq\frac{C_{p_{1}}}{R(\Delta)^{p_{1}-p\beta}}. (6.10)

Thus,

𝔼[J^1]\displaystyle\mathbb{E}[\hat{J}_{1}] CRp𝔼[|X(t)|pβ𝟏{X(t)<R1}]Cp1R(Δ)p1pβ+p=Cp1Δγ(p1pβ+p].\displaystyle\leq\frac{C}{R^{p}}\mathbb{E}\Big{[}|X(t)|^{-p\beta}\mathbf{1}_{\{X(t)<R^{-1}\}}\Big{]}\leq\frac{C_{p_{1}}}{R(\Delta)^{p_{1}-p\beta+p}}=C_{p_{1}}\Delta^{\gamma(p_{1}-p\beta+p]}. (6.11)

Similarly, we also can show that

𝔼[J^2]\displaystyle\mathbb{E}[\hat{J}_{2}] C𝔼[|X(t)|pα+p𝟏{X(t)>R}]\displaystyle\leq C\mathbb{E}\left[|X(t)|^{p\alpha+p}\mathbf{1}_{\{X(t)>R\}}\right]
Cp0R(Δ)p0p(α+1)=Cp0Δγ[p0(pα+p)].\displaystyle\leq\frac{C_{p_{0}}}{R(\Delta)^{p_{0}-p(\alpha+1)}}=C_{p_{0}}\Delta^{\gamma[p_{0}-(p\alpha+p)]}. (6.12)

By (6.9), (6.11) and (6.3), we conclude that (2.17) holds. (2.18) can be obtained in the same way as the proofs of (2.17). The proof is finished. \Box

Acknowledgment

The authors would like to thank the National Natural Science Foundation of China (12271003, 62273003, 72301173) for their financial support.

References

  • Heston [1997] S. Heston, A simple new formula for options with stochastic volatility, Washington University, St. Louis (1997).
  • Aït-Sahalia [1996] Y. Aït-Sahalia, Testing continuous-time models of the spot interest rate, Review of Financial Studies 9 (1996) 385–426.
  • Szpruch et al. [2011] L. Szpruch, X. Mao, D. J. Higham, J. Pan, Numerical simulation of a strongly nonlinear Aït-Sahalia-type interest rate model, BIT Numerical Mathematics 51 (2011) 405–425.
  • Neuenkirch and Szpruch [2014] A. Neuenkirch, L. Szpruch, First order strong approximations of scalar SDEs defined in a domain, Numerische Mathematik 128 (2014) 103–136.
  • Chassagneux et al. [2016] J.-F. Chassagneux, A. Jacquier, I. Mihaylov, An explicit Euler scheme with strong rate of convergence for financial SDEs with non-Lipschitz coefficients, SIAM Journal on Financial Mathematics 7 (2016) 993–1021.
  • Hefter and Jentzen [2018] M. Hefter, A. Jentzen, On arbitrarily slow convergence rates for strong numerical approximations of Cox-Ingersoll-Ross processes and squared Bessel processes, Finance and Stochastics 23 (2018) 139–172.
  • Zhao et al. [2020] Y. Zhao, X. Wang, M. Wang, On the backward Euler method for a generalized Ait-Sahalia-type rate model with Poisson jumps, Numerical Algorithms 87 (2020) 1321–1341.
  • Cozma and Reisinger [2020] A. Cozma, C. Reisinger, Strong order 1/2 convergence of full truncation Euler approximations to the Cox-Ingersoll-Ross process, IMA Journal of Numerical Analysis 40 (2020) 358–736.
  • Mao et al. [2021] X. Mao, F. Wei, T. Wiriyakraikul, Positivity preserving truncated Euler-Maruyama method for stochastic Lotka-Volterra competition model, Journal of Computational and Applied Mathematics 394 (2021).
  • Hong et al. [2022] J. Hong, L. Ji, X. Wang, J. Zhang, Positivity-preserving symplectic methods for the stochastic Lotka-Volterra predator-prey model, BIT Numerical Mathematics 62 (2022) 493–520.
  • Yi et al. [2021] Y. Yi, Y. Hu, J. Zhao, Positivity preserving logarithmic Euler-Maruyama type scheme for stochastic differential equations, Communications in Nonlinear Science and Numerical Simulation 101 (2021).
  • Scalone [2022] C. Scalone, Positivity preserving stochastic theta-methods for selected SDEs, Applied Numerical Mathematics 172 (2022) 351–358.
  • Lei et al. [2023] Z. Lei, S. Gan, Z. Chen, Strong and weak convergence rates of logarithmic transformed truncated EM methods for SDEs with positive solutions, Journal of Computational and Applied Mathematics 419 (2023).
  • Kelly and Lord [2023] C. Kelly, G. J. Lord, An adaptive splitting method for the Cox-Ingersoll-Ross process, Applied Numerical Mathematics 186 (2023) 252–273.
  • Chen et al. [2023] A. Chen, T. Zhou, P. Burrage, T. Tian, K. Burrage, Composite Patankar-Euler methods for positive simulations of stochastic differential equation models for biological regulatory systems, The Journal of Chemical Physics 159 (2023).
  • Hu et al. [2024] X. Hu, M. Wang, X. Dai, Y. Yu, A. Xiao, A positivity preserving Milstein-type method for stochastic differential equations with positive solutions, Journal of Computational and Applied Mathematics 449 (2024) 115963.
  • Cai et al. [2024] Y. Cai, Q. Guo, X. Mao, Strong convergence of an explicit numerical approximation for n-dimensional superlinear SDEs with positive solutions, Mathematics and Computers in Simulation 216 (2024) 198–212.
  • Jiang et al. [2024] Y. Jiang, R. Liu, X. Wang, J. Zhuo, Unconditionally positivity-preserving approximations of the Aït-Sahalia type model: explicit Milstein-type schemes, Numerical Algorithms (2024).
  • Liu et al. [2024] R. Liu, Y. Cao, X. Wang, Unconditionally positivity-preserving explicit Euler-type schemes for a generalized Aït-Sahalia model, Numerical Algorithms (2024).
  • Cai et al. [2024] Y. Cai, X. Mao, F. Wei, An advanced numerical scheme for multi-dimensional stochastic Kolmogorov equations with superlinear coefficients, Journal of Computational and Applied Mathematics 437 (2024) 115472.
  • Halidias and Stamatiou [2023] N. Halidias, I. S. Stamatiou, Boundary preserving explicit scheme for the Aït-Sahalia model, Discrete and Continuous Dynamical Systems - Series B 28 (2023) 648–664.
  • Tang and Mao [2024] Y. Tang, X. Mao, The logarithmic truncated EM method with weaker conditions, Applied Numerical Mathematics 198 (2024) 258–275.
  • Wang et al. [2020] X. Wang, J. Wu, B. Dong, Mean-square convergence rates of stochastic theta methods for SDEs under a coupled monotonicity condition, BIT Numerical Mathematics 60 (2020) 759–790.
  • Deng et al. [2023] S. Deng, C. Fei, W. Fei, X. Mao, Positivity-preserving truncated Euler-Maruyama method for generalised Aït-Sahalia-type interest model, BIT Numerical Mathematics 63 (2023).
  • Mao [2015] X. Mao, The truncated Euler-Maruyama method for stochastic differential equations, Journal of Computational and Applied Mathematics 290 (2015) 370–384.
  • Mao [2016] X. Mao, Convergence rates of the truncated Euler-Maruyama method for stochastic differential equations, Journal of Computational and Applied Mathematics 296 (2016) 362–375.
  • Szpruch et al. [2011] L. Szpruch, X. Mao, D. J. Higham, J. Pan, Numerical simulation of a strongly nonlinear Aït-Sahalia-type interest rate model, BIT Numerical Mathematics 51 (2011) 405–425.
  • Emmanuel and Mao [2021] C. Emmanuel, X. Mao, Truncated EM numerical method for generalised Ait-Sahalia-type interest rate model with delay, Journal of Computational and Applied Mathematics 383 (2021).
  • Beyn et al. [2016] W.-J. Beyn, E. Isaak, R. Kruse, Stochastic C-stability and B-consistency of explicit and implicit Euler-type schemes, Journal of Scientific Computing 67 (2016) 955–987.
  • Shi et al. [2024] B. Shi, Y. Wang, X. Mao, F. Wu, Approximation of invariant measures of a class of backward Euler-Maruyama scheme for stochastic functional differential equations, Journal of Differential Equations 389 (2024) 415–456.
  • Guo et al. [2018] Q. Guo, W. Liu, X. Mao, A note on the partially truncated Euler-Maruyama method, Applied Numerical Mathematics 130 (2018) 157–170.
  • Zhan and Li [2024] W. Zhan, Y. Li, The improvement of the truncated Euler-Maruyama method for non-Lipschitz stochastic differential equations, Advances in Computational Mathematics 50 (2024).
  • Dereich et al. [2012] S. Dereich, A. Neuenkirch, L. Szpruch, An Euler-type method for the strong approximation of the Cox-Ingersoll-Ross process, Proceedings: Mathematical Physical and Engineering Sciences 468 (2012) 1105–1115.
  • Deng et al. [2019] S. Deng, C. Fei, W. Fei, X. Mao, Generalized Aït-Sahalia-type interest rate model with Poisson jumps and convergence of the numerical approximation, Physica A: Statistical Mechanics and its Applications 533 (2019).