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

Strong approximation of some particular
one-dimensional diffusions

Madalina Deaconu1 and Samuel Herrmann2
1Université de Lorraine, CNRS, Inria, IECL, F-54000 Nancy, France,
[email protected]
2Institut de Mathématiques de Bourgogne (IMB) - UMR 5584, CNRS,
Université de Bourgogne Franche-Comté, F-21000 Dijon, France
[email protected]
Abstract

We develop a new technique for the path approximation of one-dimensional stochastic processes. Our results apply to the Brownian motion and to some families of stochastic differential equations whose distributions could be represented as a function of a time-changed Brownian motion (usually known as LL and GG-classes). We are interested here in the ε\varepsilon-strong approximation. We propose an explicit and quite easy to implement procedure that constructs jointly, the sequences of exit times and corresponding exit positions of some well chosen domains. We prove in our main results the convergence of our scheme and how to control the number of steps which depends in fact on the covering of a fixed time interval by intervals of random sizes. The underlying idea of our analysis is to combine results on Brownian exit times from time-depending domains (one-dimensional heat balls) and classical renewal theory. Numerical examples and issues are also developed in order to complete the theoretical results.

Key words: Strong approximation, path simulation, Brownian motion, linear diffusion.

2010 AMS subject classifications: primary 65C05; secondary 60J60, 60J65, 60G17.

An introduction to strong approximation

Let (Xt)t0(X_{t})_{t\geq 0} be a stochastic process defined on the filtered probability space (Ω,,(t)t,)(\Omega,{\cal{F}},({\cal{F}}_{t})_{t},\mathbb{P}) and TT be a fixed positive time. The aim of this study is to develop a new path approximation of (Xt) 0tT(X_{t})_{\ 0\leq t\leq T} where XtX_{t} stands either for the one-dimensional Brownian motion starting in xx or for a class of one-dimensional diffusions with non-homogeneous coefficients.
The usual and classical approximation procedure of any diffusion process consists in constructing numerical schemes like the Euler scheme: the time interval is split into sub-intervals 0<Tn<<n1nT<T0<\frac{T}{n}<\ldots<\frac{n-1}{n}\,T<T. For each of these time slots, the value of the process is given or approximated. The convergence result of the proposed approximation is then based on classical stochastic convergence theorems: we obtain usually some LpL^{p}-convergence between the path built by the scheme and the real path of the process. The approximation error is not a.s. bounded by a constant.
In this study, we focus our attention on a different approach: for any ε>0\varepsilon>0, we construct a suitable sequence of increasing random times (snε)n0(s_{n}^{\varepsilon})_{n\geq 0} with s0ε=0s^{\varepsilon}_{0}=0, limnsnε=+\lim_{n\to\infty}s_{n}^{\varepsilon}=+\infty on the space (Ω,,)(\Omega,{\cal{F}},\mathbb{P}), and random points x0ε,x1ε,,xnε,x_{0}^{\varepsilon},x_{1}^{\varepsilon},\ldots,x_{n}^{\varepsilon},\ldots in such a way that the random variable xnεx_{n}^{\varepsilon} is snε{\cal{F}}_{s_{n}^{\varepsilon}} adapted for any n0n\geq 0 and

supt[0,T]|Xtxtε|εa.s.\sup_{{t}\in[0,T]}|X_{t}-x^{\varepsilon}_{t}|\leq\varepsilon\quad\mbox{a.s.} (0.1)

where xtε=n0xnε1{snεt<sn+1ε}x^{\varepsilon}_{t}=\sum_{n\geq 0}x_{n}^{\varepsilon}1_{\{s_{n}^{\varepsilon}\leq t<s^{\varepsilon}_{n+1}\}}. The procedure is quite simple to describe, the sequence (snε,xnε)(s^{\varepsilon}_{n},x^{\varepsilon}_{n}) is associated to exit times and exit locations of well-chosen time-space domains for the process (t,Xt)(t,X_{t}).

We sketch the main steps of the method here. For this, let us consider a continuous function ϕε(t)\phi_{\varepsilon}(t) which satisfies: there exists rε>0r_{\varepsilon}>0 s.t.

Supp(ϕε)=[0,rε]{\rm Supp}(\phi_{\varepsilon})=[0,r_{\varepsilon}] and 0<ϕε(t)ε0<\phi_{\varepsilon}(t)\leq\varepsilon for any tSupp̊(ϕε)t\in\mathring{{\rm Supp}}(\phi_{\varepsilon}).

We start with (s0ε,x0ε)=(0,x)(s_{0}^{\varepsilon},x^{\varepsilon}_{0})=(0,x) where xx is the initial point of the path (Xt)(X_{t}). Then we define, for any n0n\geq 0

sn+1ε:=inf{tsnε:|XtXsnε|ϕε(tsnε)}s^{\varepsilon}_{n+1}:=\inf\{t\geq s^{\varepsilon}_{n}:\ |X_{t}-X_{s^{\varepsilon}_{n}}|\geq\phi_{\varepsilon}(t-s^{\varepsilon}_{n})\}

and xn+1ε:=Xsn+1εx^{\varepsilon}_{n+1}:=X_{s^{\varepsilon}_{n+1}}. In other words, sn+1εs^{\varepsilon}_{n+1} is related to the first exit time of the stochastic process (t,Xsnε+txnε)t0(t,X_{s^{\varepsilon}_{n}+t}-x^{\varepsilon}_{n})_{t\geq 0} from the time-space domain {(t,x)+×:|x|ϕε(t)}\{(t,x)\in\mathbb{R}_{+}\times\mathbb{R}:\ |x|\leq\phi_{\varepsilon}(t)\}, called ϕε\phi_{\varepsilon}-domain in the sequel.

We observe that:

  • -

    as the function ϕε\phi_{\varepsilon} is bounded, the bound (0.1) is satisfied

  • -

    as the ϕε\phi_{\varepsilon} has a compact support, the sequence (snε)(s^{\varepsilon}_{n}) satisfies sn+1εsnεrεs^{\varepsilon}_{n+1}-s^{\varepsilon}_{n}\leq r_{\varepsilon}, for any n0n\geq 0.

For such an approximation of the paths, the challenge consists in the choice of an appropriate function ϕε\phi_{\varepsilon} defining the ϕε\phi_{\varepsilon}-domain in such a way that the simulation of both the exit time and the exit location is easy to construct and implement. Moreover the analysis of the random scheme is based on a precise description of the number of random intervals [snε,sn+1ε[[s^{\varepsilon}_{n},s^{\varepsilon}_{n+1}[ required in order to cover [0,T][0,T]. Such an analysis is developed in the next section.

Our main motivation is to develop a new approach that gives the ε\varepsilon-strong approximation for a large class of multidimensional SDEs. In this paper the main tools and results of this topic are developed for some particular SDEs in one dimension. We intend to pursue this research for more general situations starting with the multidimensional Brownian motion and Bessel processes.

The study of the strong behaviour of an approximation scheme, and in particular the characterisation by some bounds depending on ε\varepsilon of supt[0,T]Xtxtε\sup_{t\in[0,T]}\|X_{t}-x^{\varepsilon}_{t}\| where xtεx^{\varepsilon}_{t} stands for an approximation scheme, was considered recently by some other authors. In Blanchet, Chen and Dong [3] the authors study the approximation of multidimensional SDEs by considering transformations of the underlying Brownian motion (the so-called Itô-Lyons map) and follow a rough path theory approach. In this paper the authors refer to the class of procedures which achieve the construction of such an approximation as Tolerance-Enforced Simulation (TES) or ε\varepsilon-strong simulation methods. In Chen and Huang [5] a similar question is considered but the result is obtained only for SDEs in one dimension and the effective construction of an approximation scheme is not obvious. This last procedure was extended by Beskos, Peluchetti and Roberts [2] were an iterative sampling method, which delivers upper and lower bounding processes for the Brownian path, is given. Let us finally mention the recent manuscript [4] which highlights an adaptation of such an approach to the fractional Brownian motion framework.

In a more general context, Hefter, Herzwurm and Müller-Gronbach [13] give lower error bounds for the pathwise approximation of scalar SDEs, the results are based on the observations of the driving Brownian motion. Previously the notion of strong convergence was studied also intensively for particular processes like the CIR process. Strong convergence without rate was obtained by Alfonsi [1] or Hutzenthaler and Jentzen [17]. Optimal lower and upper bounds were also given. For stochastic differential equations with Lipschitz coefficients Müller-Gronbach [19] and Hofmann, Müller-Gronbach and Ritter [16] obtained lower error bounds.

All these results give a new and interesting highlight in this topic of pathwise and ε\varepsilon-strong approximation, and prove how such an approach becomes an essential tool in the numerical approximation of SDEs. The procedure that we point out in this paper totally belongs to this promising field: we give an explicit and constructive procedure for the approximation of some particular SDEs. The corresponding numerical scheme is easy to implement and belongs both to the family of free knot spline approximations of scalar diffusion paths and to the family of ε\varepsilon-strong approximations. The complexity of the scheme is therefore directly linked to the number of knots required in order to describe precisely the stochastic paths on a given time-interval: Creutzig, Müller-Gronbach and Ritter [6] pointed out the smallest possible average sup-norm error depending on the average number of free knots. The important feature of the new approach of the current work is to emphasize a efficient scheme and its complexity, especially worthy for many applications. The method is essentially based on explicit distributions of the exit time for time-space domains, closely related to the behaviour of the underlying process. By performing a rigorous analysis we identify sharp estimates of the number of free knots. The analysis of the number of free knots is central in our approach.

For practical purposes the approximation scheme is the object of interest and in order to characterize and control its behaviour we are looking for sequences which have the same distribution. We need thus to introduce the following definition:

Definition 0.1. —

Let (Ω,,)(\Omega,{\cal{F}},\mathbb{P}) be a probability space and (Xt)(X_{t}) be a stochastic process on this space. The random process (ytε)(y_{t}^{\varepsilon}) is an ε\varepsilon-strong approximation of the stochastic process (Xt)(X_{t}) if there exists a stochastic process (xtε)(x_{t}^{\varepsilon}) on (Ω,,)(\Omega,{\cal{F}},\mathbb{P}) satisfying (0.1) such that (ytε)(y_{t}^{\varepsilon}) and (xtε)(x_{t}^{\varepsilon}) are identically distributed.

The material is organized as follows. In Section 1, we focus our attention on the number of space-time domains used for building the approximated path on a given fixed time interval [0,T][0,T]. This number is denoted by NTεN_{T}^{\varepsilon}. The main specific feature related to our approach is the randomness associated to the time splitting. A sharp description of the random number of time steps NTεN_{T}^{\varepsilon} permits to emphasize the efficiency of the ε\varepsilon-strong simulation. The first section points some information in a quite general framework, that is ε2𝔼[NTε]\varepsilon^{2}\mathbb{E}[N_{T}^{\varepsilon}] is upper bounded in the ε\varepsilon small limit, while the forthcoming sections permit to go into details for specific diffusion processes. Section 2 introduces the particular Brownian case and families of one-dimensional diffusions (LL-class and GG-class of diffusion in particular) are further explored in Section 3. In each case, an algorithm based on a specific ϕε\phi_{\varepsilon}-domain (heat ball) is presented (Theorem 2.1 and Theorem 3.8) and the efficiency of the approximation is investigated (Proposition 2.3 and Theorem 3.13). We obtain the convergence towards an explicit limit for the average expression ε2𝔼[NTε]\varepsilon^{2}\mathbb{E}[N_{T}^{\varepsilon}] as ε\varepsilon tends to 0. In the particular diffusion case, there exists a constant μ>0\mu>0 such that

limε0ε2𝔼[NTε]=μ𝔼[0ρ(T)1η2(x+Bs)ds],(T,x)+×\lim_{\varepsilon\to 0}\varepsilon^{2}\,\mathbb{E}[N^{\varepsilon}_{T}]=\mu\,\mathbb{E}\left[\displaystyle\int_{0}^{\rho(T)}\frac{1}{\eta^{2}(x+B_{s})}\mathrm{d}s\right],\quad\forall(T,x)\in\mathbb{R}_{+}\times\mathbb{R} (0.2)

where η\eta and ρ\rho are both functions related to the approximation procedure and (Bt)t0(B_{t})_{t\geq 0} stands for a standard one-dimensional Brownian motion.

Finally numerical examples permit to illustrate the convergence result of the algorithm in the last section.

1 Number of random intervals needed for covering the time interval [0,T][0,T]

The sharpness of the approximation is deeply related to the number of random intervals [snε,sn+1ε[[s^{\varepsilon}_{n},s^{\varepsilon}_{n+1}[ used to cover [0,T][0,T]. If (Xt)(X_{t}) is a homogeneous Markovian process, then we observe that Un+1ε=sn+1εsnεU_{n+1}^{\varepsilon}=s_{n+1}^{\varepsilon}-s_{n}^{\varepsilon}, for n0n\geq 0, is a sequence of i.i.d. a.s. bounded variables. Obviously we have:

snε=i=1nUiε.s_{n}^{\varepsilon}=\displaystyle\sum_{i=1}^{n}U_{i}^{\varepsilon}.

The number of variates corresponds to

NTε:=inf{n1:snεT}.N_{T}^{\varepsilon}:=\inf\{n\geq 1:\ s^{\varepsilon}_{n}\geq T\}.

We can control, for any jj\in\mathbb{N} and λ>0\lambda>0:

(NTε>j)=(sjε<T)=(eλsjε>eλT)eλT𝔼[eλsjε]=eλT𝔼[eλU1ε]j.\displaystyle\mathbb{P}(N_{T}^{\varepsilon}>j)=\mathbb{P}(s_{j}^{\varepsilon}<T)=\mathbb{P}(e^{-\lambda s_{j}^{\varepsilon}}>e^{-\lambda T})\leq e^{\lambda T}\mathbb{E}[e^{-\lambda s_{j}^{\varepsilon}}]=e^{\lambda T}\mathbb{E}[e^{-\lambda U_{1}^{\varepsilon}}]^{j}. (1.1)

This calculus proves that the upper-bound essentially depends on the Laplace transform of U1εU^{\varepsilon}_{1}.

Before stating a first result let us give an important convention. All along the text we need to control (upper or lower bounds) several quantities. In order to do this we use CC and κ\kappa to design positive constants, whose value may change from one line to the other. When the constants depend on parameters of prime interest, we use, for example, the notation CT,αC_{T,\alpha} to suggest that the constant CC depends in some way on TT and α\alpha, where TT and α\alpha denote here some parameters.

Proposition 1.1. —

For ε>0\varepsilon>0, let us assume that U1ε=(d)ε2UU^{\varepsilon}_{1}\stackrel{{\scriptstyle(d)}}{{=}}\varepsilon^{2}U where UU is a positive random variable which does not depend on the parameter ε\varepsilon.
1. If there exist two constants C>0C>0 and κ>0\kappa>0 such that 𝔼[eλU]Cλκ\mathbb{E}[e^{-\lambda U}]\leq\frac{C}{\lambda^{\kappa}} for all λ>0\lambda>0, then

(NTε>j)(eTC1/κjκε2)jκ,j.\mathbb{P}(N_{T}^{\varepsilon}>j)\leq\left(\frac{eTC^{1/\kappa}}{j\kappa\varepsilon^{2}}\right)^{j\kappa},\quad\forall j\in\mathbb{N}. (1.2)

2. If 𝔼[U2]<\mathbb{E}[U^{2}]<\infty, then for any δ>1\delta>1, there exists ε0>0\varepsilon_{0}>0 such that

𝔼[NTε]δeTε2𝔼[U],εε0.\mathbb{E}[N_{T}^{\varepsilon}]\leq\frac{\delta eT}{\varepsilon^{2}\mathbb{E}[U]},\quad\forall\varepsilon\leq\varepsilon_{0}.
Proof.

For the result in 1., by using both the Markov property (1.1) and the condition concerning the Laplace transform of UU, we obtain:

(NTε>j)eλT(Cε2κλκ)j,λ>0.\mathbb{P}(N_{T}^{\varepsilon}>j)\leq e^{\lambda T}\left(\frac{C}{\varepsilon^{2\kappa}\lambda^{\kappa}}\right)^{j},\quad\forall\lambda>0.

By choosing the optimal value of λ\lambda given by λ=jκT\lambda=\frac{j\kappa}{T} we obtain (1.2).
For the result in 2., we can also remark that (1.1) leads to

𝔼[NTε]=j0(NTε>j)eλT1𝔼[eλU1ε].\displaystyle\mathbb{E}[N_{T}^{\varepsilon}]=\sum_{j\geq 0}\mathbb{P}(N_{T}^{\varepsilon}>j)\leq\frac{e^{\lambda T}}{1-\mathbb{E}[e^{-\lambda U_{1}^{\varepsilon}}]}. (1.3)

If 𝔼[U2]<,\mathbb{E}[U^{2}]<\infty, we get

𝔼[eλU1ε]=1λε2𝔼[U]+o(λε2),λ>0.\mathbb{E}[e^{-\lambda U_{1}^{\varepsilon}}]=1-\lambda\varepsilon^{2}\mathbb{E}[U]+o(\lambda\varepsilon^{2}),\quad\lambda>0.

The particular choice λ=1/T\lambda=1/T implies the announced result. ∎

Remark 1.2. —

If the condition 𝔼[U2]<,\mathbb{E}[U^{2}]<\infty, is not satisfied, we can construct another approach which leads to less sharp bounds. Indeed if NTεN^{\varepsilon}_{T} denotes the number of r.v. (Unε)(U_{n}^{\varepsilon}) such that sNTεTs_{N^{\varepsilon}_{T}}\geq T, then the strong Markov property implies

𝔼[NTε]k𝔼[NT/kε],k.\mathbb{E}[N_{T}^{\varepsilon}]\leq k\mathbb{E}[N^{\varepsilon}_{T/k}],\quad\forall k\in\mathbb{N}.

Taking λ=k/T\lambda=k/T in (1.3) and afterwards k=T/ε2k=\lfloor T/\varepsilon^{2}\rfloor we obtain

𝔼[NTε]ke1𝔼[ekε2U/T]=T/ε2e1𝔼[eT/ε2ε2U/T]eTε2(1𝔼[eU])asε0.\mathbb{E}[N_{T}^{\varepsilon}]\leq\frac{ke}{1-\mathbb{E}[e^{-k\varepsilon^{2}U/T}]}=\frac{\lfloor T/\varepsilon^{2}\rfloor e}{1-\mathbb{E}[e^{-\lfloor T/\varepsilon^{2}\rfloor\varepsilon^{2}U/T}]}\sim\frac{eT}{\varepsilon^{2}(1-\mathbb{E}[e^{-U}])}\ \mbox{as}\ \varepsilon\to 0.

This result is less sharp than the statement of Proposition 1.1 since 1𝔼[eU]𝔼[U]1-\mathbb{E}[e^{-U}]\leq\mathbb{E}[U] but it holds even if the second moment of UU is not finite.

Let us just mention that the large deviations theory cannot lead to interesting bounds in our case. Indeed the rate function II used in Cramer’s theorem satisfies:

lim supnnln(NTε>n)lim supnnln(snεT)=infx[0,T]I(x)=.\limsup_{n\to\infty}n\ln\mathbb{P}(N_{T}^{\varepsilon}>n)\leq\limsup_{n\to\infty}n\ln\mathbb{P}(s^{\varepsilon}_{n}\leq T)=-\inf_{x\in[0,T]}I(x)=-\infty.

2 Approximation of one-dimensional Brownian paths

We recall that for our approach it is essential to find a function ϕε\phi_{\varepsilon} with compact support [0,rε][0,r_{\varepsilon}] which satisfies supt[0,rε]ϕε(t)=ε\sup_{t\in[0,r_{\varepsilon}]}\phi_{\varepsilon}(t)=\varepsilon and such that the exit time s1εs_{1}^{\varepsilon} of the ϕε\phi_{\varepsilon}-domain is simple to generate.

The choice of ϕε\phi_{\varepsilon} is directly related to the method of images described by Lerche [14] and to the heat equation on some particular domain, called heat-balls and defined in Evans, Section 2.3.2, [12]. More recent results on this subject can be found in [11], [10] and [9].

Brownian Skeleton (BS)η(BS)_{\eta}

  1. 1.

    Let ε>0\varepsilon>0. We define ϕε(t):=tln(ε2e/t)\phi_{\varepsilon}(t):=\sqrt{t\ln(\varepsilon^{2}e/t)}, for tIε:=[0,rε]t\in I_{\varepsilon}:=[0,r_{\varepsilon}] with rε=eε2r_{\varepsilon}=e\varepsilon^{2}.

  2. 2.

    Let (An)n1(A_{n})_{n\geq 1} be a sequence of independent random variables with gamma distribution Γ(3/2,2){\Gamma}(3/2,2)

  3. 3.

    Let (Zn)n1(Z_{n})_{n\geq 1} be a sequence of i.i.d. Rademacher random variables (taking values +1 or -1 with probability 1/2). The sequences (An)n1(A_{n})_{n\geq 1} and (Zn)n1(Z_{n})_{n\geq 1} are independent.

Definition: For ε>0\varepsilon>0 and for any function η:+\eta:\mathbb{R}\to\mathbb{R}_{+}, the Brownian skeleton (BS)η(BS)_{\eta} corresponds to

((Unε)n1,(snε)n1,(xnε)n0)with{Unε=ε2η2(xn1ε)e1An,snε=k=1nUkε,xnε=xn1ε+Znη(xn1ε)ϕε(Unεη2(xn1ε)),n1\Big{(}(U_{n}^{\varepsilon})_{n\geq 1},(s_{n}^{\varepsilon})_{n\geq 1},(x_{n}^{\varepsilon})_{n\geq 0}\Big{)}\quad\mbox{with}\quad\left\{\begin{array}[]{l}U_{n}^{\varepsilon}=\varepsilon^{2}\eta^{2}(x_{n-1}^{\varepsilon})\,e^{1-A_{n}},\quad s^{\varepsilon}_{n}=\displaystyle\sum_{k=1}^{n}U_{k}^{\varepsilon},\\[18.0pt] x_{n}^{\varepsilon}=x_{n-1}^{\varepsilon}+Z_{n}\,\eta(x_{n-1}^{\varepsilon})\phi_{\varepsilon}\left(\displaystyle\frac{U_{n}^{\varepsilon}}{\eta^{2}(x_{n-1}^{\varepsilon})}\right),\ \forall n\geq 1\\ \end{array}\right.

and x0ε=xx_{0}^{\varepsilon}=x, where xx\in\mathbb{R} fixed.

Theorem 2.1. —

Let ε>0\varepsilon>0 and let us consider a Brownian skeleton (BS)η({\rm BS})_{\eta} with η1\eta\equiv 1. Then xtε=n0xnε1{snεt<sn+1ε}x^{\varepsilon}_{t}=\sum_{n\geq 0}x_{n}^{\varepsilon}1_{\{s_{n}^{\varepsilon}\leq t<s^{\varepsilon}_{n+1}\}} is an ε\varepsilon-strong approximation of the Brownian paths starting in xx. Moreover the number of approximation points on the fixed interval [0,T][0,T] satisfies:

(NTε>j)(βTω(3/2,2,β)βjε2)j/β,j,β>2,\mathbb{P}(N_{T}^{\varepsilon}>j)\leq\left(\frac{\beta^{\prime}T\omega(3/2,2,\beta^{\prime})^{\beta^{\prime}}}{j\varepsilon^{2}}\right)^{j/\beta^{\prime}},\quad\forall j\in\mathbb{N},\ \forall\beta^{\prime}>2, (2.1)

with ω\omega a constant defined in the appendix, (4.5). Moreover, for every δ>1\delta>1 there exists ε0>0\varepsilon_{0}>0, such that the following upper-bound holds,

𝔼[NTε]33δTε2,εε0.\mathbb{E}[N_{T}^{\varepsilon}]\leq\frac{3\sqrt{3}\delta T}{\varepsilon^{2}},\quad\forall\varepsilon\leq\varepsilon_{0}.
Proof.

First we remark easily that suptIεϕε(t)=ε\sup_{t\in I_{\varepsilon}}\phi_{\varepsilon}(t)=\varepsilon as required. So we start the skeleton of the Brownian paths (BS)1({\rm BS})_{1} with the starting time-space value (0,x0ε=x)(0,x_{0}^{\varepsilon}=x). Then (0+U1ε,x0ε+Z1ϕε(U1ε))(0+U_{1}^{\varepsilon},x_{0}^{\varepsilon}+Z_{1}\phi_{\varepsilon}(U_{1}^{\varepsilon})) stands for the first exit time and exit location of the time-space domain originated in (0,x0ε)(0,x_{0}^{\varepsilon}) whose boundary is defined by ϕε\phi_{\varepsilon}.

The second step is like the first one, it suffices to consider the new starting point (s1ε,x1ε):=(U1ε,x0ε+Z1ϕε(U1ε))(s_{1}^{\varepsilon},x_{1}^{\varepsilon}):=(U_{1}^{\varepsilon},x_{0}^{\varepsilon}+Z_{1}\phi_{\varepsilon}(U_{1}^{\varepsilon})) and so on… Using the results obtained in Lerche [14] and Deaconu - Herrmann [10] (Proposition 2.2 with ν=1/2\nu=-1/2 and a=εeπ/2a=\varepsilon\sqrt{e\pi/2}), we know that these exit times are distributed like exponentials of gamma random variables (see, for instance, [10] Proposition A.2). In particular, the probability density function of U1εU_{1}^{\varepsilon} satisfies:

fU1ε(t)=ϕε(t)ε2eπt=ln(ε2e/t)ε2eπt1Iε(t),t.f_{U_{1}^{\varepsilon}}(t)=\frac{\phi_{\varepsilon}(t)}{\varepsilon\sqrt{2e\pi}t}=\frac{\sqrt{\ln(\varepsilon^{2}e/t)}}{\varepsilon\sqrt{2e\pi t}}1_{I_{\varepsilon}}(t),\quad\forall t\in\mathbb{R}.

We deduce that U1εU_{1}^{\varepsilon} and eε2We\varepsilon^{2}W defined in Lemma 4.1 are identically distributed (with the parameters α=32\alpha=\frac{3}{2} and β=2\beta=2). By Lemma 4.1, we get for any β>2\beta^{\prime}>2

𝔼[eλU1ε]ω(32,2,β)(1eε2λ)1/β.\mathbb{E}[e^{-\lambda U_{1}^{\varepsilon}}]\leq\omega\Big{(}\frac{3}{2},2,\beta^{\prime}\Big{)}\left(\frac{1}{e\varepsilon^{2}\lambda}\right)^{1/\beta^{\prime}}.

Proposition 1.1 permits to obtain the bounds of the number of points needed to approximate the Brownian paths on the interval [0,T][0,T], as 𝔼(W)=133\mathbb{E}(W)=\frac{1}{3\sqrt{3}}. ∎

Remark 2.2. —
  1. 1.

    Let us just notice that for UU a standard uniformly distributed r.v. and GG a standard Gaussian r.v. independent of UU, W=U2eG2W=U^{2}e^{-G^{2}} is random variable with the PDF presented in Lemma 4.1 associated to the parameters α=32\alpha=\frac{3}{2} and β=2\beta=2 (for more details see [8], Chapter IX.3).

  2. 2.

    A similar approach will be used in the proof of Theorem 3.8. In Theorem 2.1 we obtain that for η1\eta\equiv 1, we can construct a sequence of successive points corresponding to the exit time and location of ε\varepsilon-small spheroids and belonging to the Brownian trajectory. Moreover this sequence has the same distribution as (Unε,xnε)n1(U_{n}^{\varepsilon},x_{n}^{\varepsilon})_{n\geq 1}. This procedure can also be considered for general η\eta: Un+1εη2(xnε)\frac{U_{n+1}^{\varepsilon}}{\eta^{2}(x_{n}^{\varepsilon})} has then the same distribution as the Brownian first exit time of the ε\varepsilon-small spheroids. Therefore for any t[snε,sn+1ε]t\in[s_{n}^{\varepsilon},s_{n+1}^{\varepsilon}], we get

    |x0ε+Btxnε|εη(xnε),|x_{0}^{\varepsilon}+B_{t}-x_{n}^{\varepsilon}|\leq\varepsilon\eta(x_{n}^{\varepsilon}),

    where BB stands for the standard Brownian motion.

We can easily improve the description of the number of approximation points. Since (Unε)n0(U_{n}^{\varepsilon})_{n\geq 0} is a sequence of independent random variables, (Ntε)t0(N^{\varepsilon}_{t})_{t\geq 0} is a renewal process and the classical asymptotic description holds:

Proposition 2.3. —

We consider the Brownian skeleton (BS)η(BS)_{\eta} for η=1\eta=1. We define, as previously

NTε:=inf{n1:snεT}.N_{T}^{\varepsilon}:=\inf\{n\geq 1:\ s^{\varepsilon}_{n}\geq T\}.

the number of approximation points needed to cover the time interval [0,T][0,T] for TT a fixed positive time. Then:

limε0ε2𝔼[NTε]=Te 33/2.\lim_{\varepsilon\to 0}\varepsilon^{2}\mathbb{E}[N^{\varepsilon}_{T}]=\frac{T}{e}\,3^{3/2}.

Moreover the following central limit theorem holds:

limε0μ3ε2σ2T(ε2NTεTe 33/2)=Gin distribution,\displaystyle\lim_{\varepsilon\to 0}\sqrt{\frac{\mu^{3}}{\varepsilon^{2}\sigma^{2}T}}\Big{(}\varepsilon^{2}N^{\varepsilon}_{T}-\frac{T}{e}\,3^{3/2}\Big{)}=G\quad\mbox{in distribution},

where GG is a 𝒩(0,1)\mathcal{N}(0,1) standard Gaussian random variable, μ=e 33/20.5231336\mu=e\,3^{-3/2}\approx 0.5231336 and σ2=(53/233)e20.3872285\sigma^{2}=(5^{-3/2}-3^{-3})\,e^{2}\approx 0.3872285.

Proof.

Let us consider (N¯t)t0(\overline{N}_{t})_{t\geq 0} a renewal process with interarrivals (e1An)n1(e^{1-A_{n}})_{n\geq 1} independent random variables defined in Theorem 2.1. The interarrival time satisfies 𝔼[e1A1]=eA1(1)\mathbb{E}[e^{1-A_{1}}]=e\mathcal{L}_{A_{1}}(1) where A1\mathcal{L}_{A_{1}} stands for the Laplace transform of A1A_{1}. Since in our case A1A_{1} is gamma distributed, it is well known that A1(s)=(2s+1)3/2\mathcal{L}_{A_{1}}(s)=(2s+1)^{-3/2}.
We use here classical results for the renewal theory, see for example [7]. The elementary renewal theorem leads to

limt𝔼[N¯t]t=1𝔼[e1A1]=33/2e.\lim_{t\to\infty}\frac{\mathbb{E}[\overline{N}_{t}]}{t}=\frac{1}{\mathbb{E}[e^{1-A_{1}}]}=\frac{3^{3/2}}{e}.

In order to obtain the first part of the statement, it suffices to observe that:

NTε\displaystyle N_{T}^{\varepsilon} =inf{n0:k=1ne1AkTε2}=N¯T/ε2.\displaystyle=\inf\Big{\{}n\geq 0:\ \sum_{k=1}^{n}e^{1-A_{k}}\geq\frac{T}{\varepsilon^{2}}\Big{\}}=\overline{N}_{T/\varepsilon^{2}}.

We deduce that

limε0ε2𝔼[NTε]=limε0ε2𝔼[N¯T/ε2]=limtT𝔼[N¯t]t=Te 33/2.\lim_{\varepsilon\to 0}\varepsilon^{2}\mathbb{E}[N_{T}^{\varepsilon}]=\lim_{\varepsilon\to 0}\varepsilon^{2}\mathbb{E}[\overline{N}_{T/\varepsilon^{2}}]=\lim_{t\to\infty}T\frac{\mathbb{E}[\overline{N}_{t}]}{t}=\frac{T}{e}\,3^{3/2}.

The same argument holds for the CLT: if we denote by μ=𝔼[e1A1]\mu=\mathbb{E}[e^{1-A_{1}}] and σ2=Var(e1A1)\sigma^{2}={\rm Var}(e^{1-A_{1}}) then

limε0tμ3σ2(N¯tt1μ)=Gin distribution,\displaystyle\lim_{\varepsilon\to 0}\sqrt{\frac{t\mu^{3}}{\sigma^{2}}}\Big{(}\frac{\overline{N}_{t}}{t}-\frac{1}{\mu}\Big{)}=G\quad\mbox{in distribution},

where GG is a 𝒩(0,1)\mathcal{N}(0,1) standard Gaussian random variable. The statement is therefore a consequence of the link between NTεN^{\varepsilon}_{T} and N¯T/ε2\overline{N}_{T/\varepsilon^{2}}. ∎

3 The particular L and G classes of diffusion

Let us now consider some generalizations of the Brownian paths study. We introduce solutions of the following one-dimensional stochastic differential equation:

dXt=σ(t,Xt)dBt+μ(t,Xt)dt,X0=x0,dX_{t}=\sigma(t,X_{t})dB_{t}+\mu(t,X_{t})\,dt,\quad X_{0}=x_{0}, (3.1)

where (Bt,t0)(B_{t},\ t\geq 0) stands for a standard one-dimensional Brownian motion and σ,μ:[0,+)×\sigma,\mu:[0,+\infty)\times\mathbb{R}\to\mathbb{R}. Let us consider two families of diffusions introduced in Wang - Pötzelberger [20]:

  1. 1.

    (LL-class) for σ(t,x)=σ¯(t)\sigma(t,x)=\overline{\sigma}(t) and μ(t,x)=a(t)x+b(t)\mu(t,x)=a(t)x+b(t), xx\in\mathbb{R}

  2. 2.

    (GG-class) for σ(t,x)=σ¯x\sigma(t,x)=\underline{\sigma}x and μ(t,x)=a(t)x+b(t)xln(x)\mu(t,x)=a(t)x+b(t)x\ln(x), x+x\in\mathbb{R}_{+},

where σ¯:++,a,b:+\overline{\sigma}:\mathbb{R}_{+}\to\mathbb{R}_{+},a,b:\mathbb{R}_{+}\to\mathbb{R} are 𝒞1\mathcal{C}^{1}-functions and σ¯+\underline{\sigma}\in\mathbb{R}_{+}.

Let us note that, in such particular cases, the solution of the SDE (3.1) has the same distribution as a function of the time-changed Brownian motion:

Xt=f(t,x0+Bρ(t)),t0,X_{t}=f(t,x_{0}+B_{\rho(t)}),\quad t\geq 0, (3.2)

(where ff and ρ\rho denote functions that we specify for each class afterwards).

For LL-class diffusions for instance one particular choice of the function ff (this choice is not unique) is given by (see, for instance, Karatzas and Shreve [18], p. 354, Section 5.6 for classical formulas and Herrmann and Massin [15] for new developments in this topic):

f(t,x)=σ¯(t)ρ(t)x+c(t),f(t,x)=\frac{\overline{\sigma}(t)}{\sqrt{\rho^{\prime}(t)}}\,x+c(t), (3.3)

with

c(t)=e0ta(s)𝑑s0tb(s)e0sa(u)𝑑u𝑑s,andρ(t)=0tσ¯2(s)e20sa(u)𝑑u𝑑s.c(t)=e^{\int_{0}^{t}a(s)\,ds}\int_{0}^{t}b(s)e^{-\int_{0}^{s}a(u)\,du}\,ds,\quad\mbox{and}\quad\rho(t)=\int_{0}^{t}\overline{\sigma}^{2}(s)e^{-2\int_{0}^{s}a(u)\,du}\,ds.
Remark 3.1. —

If we have a diffusion in the LL-class characterized by some fixed function f(t,x)f(t,x) given in (3.2) then we can obtain a diffusion of GG-class by using the function ef(t,x)e^{f(t,x)} instead of f(t,x)f(t,x). Obviously the corresponding coefficients a,b,a,b, and σ¯\underline{\sigma} need to be specified with respect to those connected to f(t,x)f(t,x).

Proposition 3.2. —

Let us define the following diffusion process

Xt=f(t,x0+Bρ(t)),t0,X_{t}=f(t,x_{0}+B_{\rho(t)}),\quad t\geq 0, (3.4)

where ff is given by (3.3). Then XtX_{t} is a weak solution of the stochastic differential equation (3.1).

Proof.

We can write the previous expression for ff on the form

f(t,x)=xe0ta(s)ds+e0ta(s)ds0tb(s)e0sa(u)duds.f(t,x)=x\cdot e^{\int_{0}^{t}a(s)\mathrm{d}s}+e^{\int_{0}^{t}a(s)\mathrm{d}s}\displaystyle\int_{0}^{t}b(s)e^{-\int_{0}^{s}a(u)\mathrm{d}u}\mathrm{d}s. (3.5)

We denote also by

𝕀t=0tρ(s)dBs.\mathbb{I}_{t}=\displaystyle\int_{0}^{t}\sqrt{\rho^{\prime}(s)}\mathrm{d}B_{s}. (3.6)

In order to prove the result we need to prove that Xt=f(t,x0+𝕀t)X_{t}=f(t,x_{0}+\mathbb{I}_{t}) satisfies the equation (3.1). For the initial condition we can see that:

X0=f(0,x0)=x0.X_{0}=f(0,x_{0})=x_{0}. (3.7)

Let us now evaluate

dXt=[a(t)e0ta(s)ds(x0+𝕀t)+a(t)e0ta(s)ds0tb(s)e0sa(u)duds+e0ta(s)dsb(t)e0ta(s)ds]dt+σ¯(t)dBt=[a(t)f(t,x0+𝕀t)+b(t)]dt+σ¯(t)dBt=[a(t)Xt+b(t)]dt+σ¯dBt,\begin{array}[]{ll}\mathrm{d}X_{t}&=\left[a(t)e^{\int_{0}^{t}a(s)\mathrm{d}s}(x_{0}+\mathbb{I}_{t})+a(t)e^{\int_{0}^{t}a(s)\mathrm{d}s}\displaystyle\int_{0}^{t}b(s)e^{-\int_{0}^{s}a(u)\mathrm{d}u}\mathrm{d}s+e^{\int_{0}^{t}a(s)\mathrm{d}s}b(t)e^{-\int_{0}^{t}a(s)\mathrm{d}s}\right]\mathrm{d}t\\ &\qquad+\overline{\sigma}(t)\mathrm{d}B_{t}\\ &=\left[a(t)f(t,x_{0}+\mathbb{I}_{t})+b(t)\right]\mathrm{d}t+\overline{\sigma}(t)\mathrm{d}B_{t}=\left[a(t)X_{t}+b(t)\right]\mathrm{d}t+\overline{\sigma}\mathrm{d}B_{t},\end{array} (3.8)

by using Itô formula.
This ends the proof of the proposition. ∎

In this section, we consider particular diffusion processes which are strongly related to the Brownian paths. It is therefore intuitive to replace in (3.2) the Brownian trajectory by its approximation. If the function ff is Lipschitz continuous, then the error stemmed from the approximation is easily controlled (the proof is left to the reader).

Assumption 3.3. —

The diffusion process (Xt)t0(X_{t})_{t\geq 0} satisfies

Xt=f(t,x0+Bρ(t))X_{t}=f(t,x_{0}+B_{\rho(t)})

with ff a Lipschitz continuous function:

|f(t,x)f(s,y)|KLip(T)(|xy|+|ts|),(x,y)2,(s,t)[0,T]2|f(t,x)-f(s,y)|\leq K_{Lip}(T)(|x-y|+|t-s|),\quad\forall(x,y)\in\mathbb{R}^{2},\quad\forall(s,t)\in[0,T]^{2} (3.9)

where KLip(T)K_{Lip}(T) stands for the Lipschitz constant. The function ρ\rho is a strictly increasing continuous function with initial value ρ(0)=0\rho(0)=0.

Proposition 3.4. —

Consider T>0T>0 and ε>0\varepsilon>0 fixed. Let the diffusion process (Xt)(X_{t}) satisfy Assumption 3.3 and let

xtθ:=n0xnθ1{snθt<sn+1θ}x_{t}^{\theta}:=\sum_{n\geq 0}x_{n}^{\theta}1_{\{s_{n}^{\theta}\leq t<s_{n+1}^{\theta}\}}

be a θ\theta-strong approximation of the Brownian motion (see Theorem 2.1) with θ=εKLip1(ρ1(T))\theta=\varepsilon K_{Lip}^{-1}(\rho^{-1}(T)) on the time interval [0,ρ1(T)][0,\rho^{-1}(T)], where KLip(T)K_{Lip}(T) is defined in (3.9), then

ytε:=n0f(ρ1(snθ),xnθ)1{snθρ(t)<sn+1θ}y^{\varepsilon}_{t}:=\sum_{n\geq 0}f(\rho^{-1}(s_{n}^{\theta}),x_{n}^{\theta})1_{\{s_{n}^{\theta}\leq\rho(t)<s_{n+1}^{\theta}\}}

is an ε\varepsilon-strong approximation of (Xt)(X_{t}) on [0,T][0,T].

Unfortunately the Lipschitz continuity of the function ff is a restrictive condition which is not relevant for most of the diffusion processes. In particular, a typical diffusion belonging to the LL or GG-class does not satisfy the Lipschitz condition. Consequently we introduce a more general framework.

Assumption 3.5. —

The diffusion process (Xt)t0(X_{t})_{t\geq 0} is a function of the time-changed Brownian motion:

Xt=f(t,x0+Bρ(t))X_{t}=f(t,x_{0}+B_{\rho(t)})

where ρ\rho is an increasing differentiable function satisfying limtρ(t)=\lim_{t\to\infty}\rho(t)=\infty with initial value ρ(0)=0\rho(0)=0, ff is a 𝒞1,1(+×,)\mathcal{C}^{1,1}(\mathbb{R}_{+}\times\mathbb{R},\mathbb{R})-function. There exists a strictly increasing 𝒞2()\mathcal{C}^{2}(\mathbb{R})-function FF such that

supt[0,T]max{|ft(t,x)|,|fx(t,x)|}F(x2),x.\sup_{t\in[0,T]}\max\Big{\{}\left|\frac{\partial f}{\partial t}(t,x)\right|,\left|\frac{\partial f}{\partial x}(t,x)\right|\Big{\}}\leq F(x^{2}),\quad\forall x\in\mathbb{R}. (3.10)

Furthermore we assume:

  • -

    there exist two constants κ1\kappa_{1} and κ2\kappa_{2} such that max(F,F,F′′)(x2)κ1eκ2x\max(F,F^{\prime},F^{\prime\prime})(x^{2})\leq\kappa_{1}e^{\kappa_{2}x} for all x0x\geq 0

  • -

    the function x1F(2x2+1)x\mapsto\frac{1}{F(2x^{2}+1)} is Lipschitz-continuous.

Assumption 3.6. —

κmin>0\exists\kappa_{\min}>0 such that ρ(ρ1(x))κmin\rho^{\prime}(\rho^{-1}(x))\geq\kappa_{\rm min} for all xx\in\mathbb{R}.

Remark 3.7. —

One can check easily that the L and GG-class diffusions verify these hypotheses.

Let us define the function η\eta to be

η(x)=1(eκmin1+1)F(2x2+1).\eta(x)=\frac{1}{(e\kappa_{\rm min}^{-1}+1)F(2x^{2}+1)}. (3.11)

By Assumption 3.5 this function is strictly decreasing and Lipschitz continuous.

Theorem 3.8. —

Let ε>0\varepsilon>0. Let (Xt)t0(X_{t})_{t\geq 0} be the solution of (3.1) satisfying Assumptions 3.5 and 3.6 and let us consider the Brownian skeleton (BS)η({\rm BS})_{\eta} associated to the function η\eta defined in (3.11), then

ytε:=n0f(ρ1(snε),xnε)1{snερ(t)<sn+1ε}y^{\varepsilon}_{t}:=\sum_{n\geq 0}f(\rho^{-1}(s_{n}^{\varepsilon}),x_{n}^{\varepsilon})1_{\{s_{n}^{\varepsilon}\leq\rho(t)<s_{n+1}^{\varepsilon}\}} (3.12)

is an ε\varepsilon-strong approximation of (Xt)(X_{t}) on [0,T][0,T].

Proof.

Let us assume that tt satisfies snερ(t)sn+1εs_{n}^{\varepsilon}\leq\rho(t)\leq s_{n+1}^{\varepsilon} for some nn\in\mathbb{N}. We denote tnε:=ρ1(snε)t_{n}^{\varepsilon}:=\rho^{-1}(s_{n}^{\varepsilon}) and Atn:=f(t,x0ε+Bρ(t))f(tnε,xnε)A^{n}_{t}:=f(t,x_{0}^{\varepsilon}+B_{\rho(t)})-f(t_{n}^{\varepsilon},x_{n}^{\varepsilon}). We obtain, there exists τ(0,1)\tau\in(0,1) such that:

Atn\displaystyle A^{n}_{t} =(ttnε)ft(tnε+τ(ttnε),xnε+τ(x0ε+Bρ(t)xnε))\displaystyle=(t-t_{n}^{\varepsilon})\frac{\partial f}{\partial t}(t_{n}^{\varepsilon}+\tau(t-t_{n}^{\varepsilon}),x_{n}^{\varepsilon}+\tau(x_{0}^{\varepsilon}+B_{\rho(t)}-x_{n}^{\varepsilon}))
+(x0ε+Bρ(t)xnε)fx(tnε+τ(ttnε),xnε+τ(x0ε+Bρ(t)xnε)).\displaystyle\quad+(x_{0}^{\varepsilon}+B_{\rho(t)}-x_{n}^{\varepsilon})\frac{\partial f}{\partial x}(t_{n}^{\varepsilon}+\tau(t-t_{n}^{\varepsilon}),x_{n}^{\varepsilon}+\tau(x_{0}^{\varepsilon}+B_{\rho(t)}-x_{n}^{\varepsilon})).

Under the assumption (3.10) we have

|Atn|\displaystyle|A^{n}_{t}| (|ttnε|+|x0ε+Bρ(t)xnε|)F((xnε+τ(x0ε+Bρ(t)xnε))2).\displaystyle\leq\Big{(}|t-t_{n}^{\varepsilon}|+|x_{0}^{\varepsilon}+B_{\rho(t)}-x_{n}^{\varepsilon}|\Big{)}\cdot F((x_{n}^{\varepsilon}+\tau(x_{0}^{\varepsilon}+B_{\rho(t)}-x_{n}^{\varepsilon}))^{2}).

Since

|ttnε||tn+1εtnε|=|ρ1(sn+1ε)ρ1(snε)|=|snεsn+1εduρ(ρ1(u))|,|t-t_{n}^{\varepsilon}|\leq|t_{n+1}^{\varepsilon}-t_{n}^{\varepsilon}|=|\rho^{-1}(s_{n+1}^{\varepsilon})-\rho^{-1}(s_{n}^{\varepsilon})|=\Big{|}\int_{s_{n}^{\varepsilon}}^{s_{n+1}^{\varepsilon}}\frac{\mathrm{d}u}{\rho^{\prime}(\rho^{-1}(u))}\Big{|},

we obtain

|ttnε|κmin1|sn+1εsnε|=κmin1|Un+1ε|eκmin1ε2η2(xnε).|t-t_{n}^{\varepsilon}|\leq\kappa_{\rm min}^{-1}|s_{n+1}^{\varepsilon}-s_{n}^{\varepsilon}|=\kappa_{\rm min}^{-1}|U_{n+1}^{\varepsilon}|\leq e\,\kappa_{\rm min}^{-1}\varepsilon^{2}\eta^{2}(x_{n}^{\varepsilon}).

Moreover, by the definition of the BM approximation (see Remark 2.2),

|x0ε+Bρ(t)xnε|εη(xnε).|x_{0}^{\varepsilon}+B_{\rho(t)}-x_{n}^{\varepsilon}|\leq\varepsilon\eta(x_{n}^{\varepsilon}).

Finally due to the monotone property of FF,

|Atn|(eκmin1ε2η2(xnε)+εη(xnε))F((xnε+τ(x0ε+Bρ(t)xnε))2)(eκmin1ε2η2(xnε)+εη(xnε))F(2(xnε)2+2ε2η2(xnε)).\begin{array}[]{ll}|A^{n}_{t}|&\leq\Big{(}e\,\kappa_{\rm min}^{-1}\varepsilon^{2}\eta^{2}(x_{n}^{\varepsilon})+\varepsilon\eta(x_{n}^{\varepsilon})\Big{)}\cdot F((x_{n}^{\varepsilon}+\tau(x_{0}^{\varepsilon}+B_{\rho(t)}-x_{n}^{\varepsilon}))^{2})\\ &\leq\Big{(}e\,\kappa_{\rm min}^{-1}\varepsilon^{2}\eta^{2}(x_{n}^{\varepsilon})+\varepsilon\eta(x_{n}^{\varepsilon})\Big{)}\cdot F(2(x_{n}^{\varepsilon})^{2}+2\varepsilon^{2}\eta^{2}(x_{n}^{\varepsilon})).\\ \end{array}

There exists ε0>0\varepsilon_{0}>0 such that ε2η2(x)1/2\varepsilon^{2}\eta^{2}(x)\leq 1/2 for all xx\in\mathbb{R} and εε0\varepsilon\leq\varepsilon_{0}. Then, by the definition of the function η\eta, for εε0\varepsilon\leq\varepsilon_{0}, we have

|Atn|(eκmin1+1)εη(xnε)F(2(xnε)2+1)ε,|A^{n}_{t}|\leq(e\,\kappa_{\rm min}^{-1}+1)\varepsilon\eta(x_{n}^{\varepsilon})\cdot F(2(x_{n}^{\varepsilon})^{2}+1)\leq\varepsilon,

for any t[snε,sn+1ε]t\in[s_{n}^{\varepsilon},s_{n+1}^{\varepsilon}]. We deduce that the piecewise constant approximation associated to (ynε)n(y_{n}^{\varepsilon})_{n} where ynε:=f(ρ1(snε),xnε)y_{n}^{\varepsilon}:=f(\rho^{-1}(s_{n}^{\varepsilon}),x_{n}^{\varepsilon}) and ninf{k0:skρ(T)}n\leq\inf\{k\geq 0:\ s_{k}\geq\rho(T)\} is a ε\varepsilon-strong approximation of (Xt,t[0,T])(X_{t},\,t\in[0,T]). ∎

Let us now describe the efficiency of the ε\varepsilon-strong approximation. We introduce

Ntε:=inf{n0:snεt}andN^tε:=Nρ(t)ε,N_{t}^{\varepsilon}:=\inf\{n\geq 0:\ s_{n}^{\varepsilon}\geq t\}\quad\mbox{and}\quad\hat{N}_{t}^{\varepsilon}:=N_{\rho(t)}^{\varepsilon}, (3.13)

where snεs_{n}^{\varepsilon} is issued from the Brownian skeleton (BS)η({\rm BS})_{\eta}. N^tε\hat{N}_{t}^{\varepsilon} corresponds therefore to the number of random points needed to approximate the diffusion paths on [0,t][0,t]. Let us observe that the random variables UnεU_{n}^{\varepsilon} are no more i.i.d. random variables in the diffusion case (different to the Brownian case), therefore we cannot use the classical renewal theorem in order to describe N^tε\hat{N}_{t}^{\varepsilon}.

Proposition 3.9. —

Let (Ntε,t0)(N^{\varepsilon}_{t},\,t\geq 0) be the counting process defined by (3.13). Then, under Assumptions 3.5 and 3.6, for any xx\in\mathbb{R} and t0t\geq 0, the average ψε(t,x):=𝔼[Ntε|x0ε=x]\psi^{\varepsilon}(t,x):=\mathbb{E}[N^{\varepsilon}_{t}|x_{0}^{\varepsilon}=x] is finite. Moreover there exists a constant λ0>0\lambda_{0}>0 such that the Laplace transform ψε(λ,x):=0eλtψε(t,x)𝑑t\mathcal{L}\psi^{\varepsilon}(\lambda,x):=\int_{0}^{\infty}e^{-\lambda t}\psi^{\varepsilon}(t,x)\,dt is finite for any λ\lambda\in\mathbb{C} satisfying Re(λ)>λ0{\rm Re}(\lambda)>\lambda_{0}.

Proof.

The mean of the counting process is defined by

ψε(t,x)=n1(Ntεn)=n1(snεt).\psi^{\varepsilon}(t,x)=\sum_{n\geq 1}\mathbb{P}(N_{t}^{\varepsilon}\geq n)=\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t).

This equality holds since snεs_{n}^{\varepsilon} is a continuous random variables by the definition of (BS)η({\rm BS})_{\eta}. Let us denote by mnε:=min0kn1η2(xkε)m_{n}^{\varepsilon}:=\min_{0\leq k\leq n-1}\eta^{2}(x_{k}^{\varepsilon}), where η\eta is defined by (3.11) and introduce the following decomposition:

(snεt)=(snεt,mnεn2/3)+(snεt,mnε<n2/3).\mathbb{P}(s_{n}^{\varepsilon}\leq t)=\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,m_{n}^{\varepsilon}\geq n^{-2/3})+\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,m_{n}^{\varepsilon}<n^{-2/3}). (3.14)

By the definition of the sequence (snε)(s_{n}^{\varepsilon}), we get

snε=ε2η2(x0ε)e1A1++ε2η2(xn1ε)e1Anε2mnε(e1A1++e1An).s_{n}^{\varepsilon}=\varepsilon^{2}\eta^{2}(x_{0}^{\varepsilon})e^{1-A_{1}}+\ldots+\varepsilon^{2}\eta^{2}(x_{n-1}^{\varepsilon})\,e^{1-A_{n}}\geq\varepsilon^{2}m_{n}^{\varepsilon}(e^{1-A_{1}}+\ldots+e^{1-A_{n}}).

Hence, for any λ>0\lambda>0, we have

unε(t):=(snεt,mnεn2/3)\displaystyle u_{n}^{\varepsilon}(t):=\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,m_{n}^{\varepsilon}\geq n^{-2/3}) (e1A1++e1Antn2/3ε2)\displaystyle\leq\mathbb{P}(e^{1-A_{1}}+\ldots+e^{1-A_{n}}\leq tn^{2/3}\varepsilon^{-2})
=(exp{λ2n2/3ε2(e1A1++e1An)}eλt2)\displaystyle=\mathbb{P}\left(\exp\left\{-\frac{\lambda}{2}n^{-2/3}\varepsilon^{2}(e^{1-A_{1}}+\ldots+e^{1-A_{n}})\right\}\geq e^{-\frac{\lambda t}{2}}\right)
eλt2𝔼[exp(λ2n2/3ε2e1A1)]n.\displaystyle\leq e^{\frac{\lambda t}{2}}\mathbb{E}\Big{[}\exp\left(-\frac{\lambda}{2}\,n^{-2/3}\varepsilon^{2}e^{1-A_{1}}\right)\Big{]}^{n}. (3.15)

Since the second moment of e1A1e^{1-A_{1}} is finite, we obtain the Taylor expansion:

𝔼[exp(λ2n2/3ε2e1A1)]=1λ2n2/3ε2e33/2+(λ2)2n4/3ε4e2253/2+o(n4/3).\mathbb{E}\left[\exp\left(-\frac{\lambda}{2}n^{-2/3}\varepsilon^{2}e^{1-A_{1}}\right)\right]=1-\frac{\lambda}{2}\frac{n^{-2/3}\varepsilon^{2}e}{3^{3/2}}+\left(\frac{\lambda}{2}\right)^{2}\frac{n^{-4/3}\varepsilon^{4}e^{2}}{2\cdot 5^{3/2}}+o(n^{-4/3}).

By using the classical relation ln(1x)=(x+x22+x33+)\ln(1-x)=-\left(x+\frac{x^{2}}{2}+\frac{x^{3}}{3}+\ldots\right), for x(0,1)x\in(0,1), we can deduce that

𝔼[exp(λ2n2/3ε2e1A1)]n=exp[λ2en1/3ε233/2]+o(1) as n+.\mathbb{E}\left[\exp\left(-\displaystyle\frac{\lambda}{2}n^{-2/3}\varepsilon^{2}e^{1-A_{1}}\right)\right]^{n}=\exp\left[-\displaystyle\frac{\lambda}{2}\displaystyle\frac{en^{1/3}\varepsilon^{2}}{3^{3/2}}\right]+o(1)\mbox{ as }n\to+\infty. (3.16)

Using classical results on series with positive terms, we obtain by comparison

n1(snεt,mnεn2/3)<.\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,m_{n}^{\varepsilon}\geq n^{-2/3})<\infty. (3.17)

Let us just note that this result is still true if we consider the terms u¯nε(λ):=0eλtunε(t)𝑑t\overline{u}_{n}^{\varepsilon}(\lambda):=\int_{0}^{\infty}e^{-\lambda t}u_{n}^{\varepsilon}(t)\,dt. Indeed (3) leads to

u¯nε(λ)0eλteλt2𝔼[exp(λ2n2/3ε2e1A1)]ndt=2λ𝔼[exp(λ2n2/3ε2e1A1)]n.\overline{u}_{n}^{\varepsilon}(\lambda)\leq\int_{0}^{\infty}e^{-\lambda t}e^{\frac{\lambda t}{2}}\mathbb{E}\left[\exp\left(-\displaystyle\frac{\lambda}{2}n^{-2/3}\varepsilon^{2}e^{1-A_{1}}\right)\right]^{n}\,\mathrm{d}t=\frac{2}{\lambda}\,\mathbb{E}\left[\exp\left(-\displaystyle\frac{\lambda}{2}n^{-2/3}\varepsilon^{2}e^{1-A_{1}}\right)\right]^{n}.

Since the upper bound is the term of a convergent series, we deduce by comparison that

n1u¯nε(λ)<,λ>0.\sum_{n\geq 1}\overline{u}_{n}^{\varepsilon}(\lambda)<\infty,\quad\forall\lambda>0. (3.18)

Let us now focus our attention to the second term of the r.h.s in (3.14). Since we consider the Brownian skeleton (BS)η({\rm BS})_{\eta}, the sequence (snε,xnε)(s_{n}^{\varepsilon},x_{n}^{\varepsilon}) belongs to the graph of a Brownian trajectory (see Remark 2.2). Consequently the condition mnε<n2/3m_{n}^{\varepsilon}<n^{-2/3} can be related to a condition on the Brownian paths:

vnε(t):=(snεt,mnε<n2/3)(sts.t.η2(x0ε+Bs)<n2/3),v_{n}^{\varepsilon}(t):=\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,m_{n}^{\varepsilon}<n^{-2/3})\leq\mathbb{P}(\exists s\leq t\ {\rm s.t.}\ \eta^{2}(x_{0}^{\varepsilon}+B_{s})<n^{-2/3}),

where BB is a one-dimensional standard Brownian motion. Using the upper-bound of the function FF in Assumption 3.5 and the definition of the function η\eta, we obtain the bound: there exists C>0C>0 and κ>0\kappa>0 such that η(x)Ceκ|x|\eta(x)\geq Ce^{-\kappa|x|}. Let us note that x0ε=xx_{0}^{\varepsilon}=x. The Brownian reflection principle leads to

vnε(t)\displaystyle v_{n}^{\varepsilon}(t) (η2(sups[0,t]|x+Bs|)<n2/3)\displaystyle\leq\mathbb{P}\left(\eta^{2}\left(\sup_{s\in[0,t]}|x+B_{s}|\right)<n^{-2/3}\right)
(sups[0,t]|x+Bs|>13κln(nC3))\displaystyle\leq\mathbb{P}\Big{(}\sup_{s\in[0,t]}|x+B_{s}|>\frac{1}{3\kappa}\,\ln(nC^{3})\Big{)}
(sups[0,t]|Bs|>13κln(nC3)|x|)\displaystyle\leq\mathbb{P}\Big{(}\sup_{s\in[0,t]}|B_{s}|>\frac{1}{3\kappa}\,\ln(nC^{3})-|x|\Big{)}
2(|Bt|>13κln(nC3)|x|)\displaystyle\leq 2\mathbb{P}\Big{(}|B_{t}|>\frac{1}{3\kappa}\,\ln(nC^{3})-|x|\Big{)}
12κtln(nC3/e3κ|x|)2πexp(ln2(nC3/e3κ|x|)72κ2t).\displaystyle\leq\frac{12\kappa\sqrt{t}}{\ln(nC^{3}/e^{3\kappa|x|})\sqrt{2\pi}}\,\exp\Big{(}-\frac{\ln^{2}(nC^{3}/e^{3\kappa|x|})}{72\kappa^{2}t}\Big{)}.

Since ln2(nC3/e3κ|x|)72κ2t2ln(n)\frac{\ln^{2}(nC^{3}/e^{3\kappa|x|})}{72\kappa^{2}t}\geq 2\ln(n) for large values of nn, we deduce that the r.h.s of the previous equality corresponds to the term of a convergent series. Therefore

n1(snεt,mnε<n2/3)<.\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,m_{n}^{\varepsilon}<n^{-2/3})<\infty. (3.19)

This finishes the proof of finiteness of ψε(t,x)\psi^{\varepsilon}(t,x).
Let us now define v¯nε(λ):=0eλtvnε(t)dt\overline{v}_{n}^{\varepsilon}(\lambda):=\int_{0}^{\infty}e^{-\lambda t}v_{n}^{\varepsilon}(t)\,\mathrm{d}t. The previous inequalities permit to obtain:

v¯nε(λ)0eλt2(|Bt|>13κln(nC3/e3κ|x|))dt=2π+2eλtu221{uαn/t}dtdu\displaystyle\overline{v}_{n}^{\varepsilon}(\lambda)\leq\int_{0}^{\infty}e^{-\lambda t}2\mathbb{P}\Big{(}|B_{t}|>\frac{1}{3\kappa}\,\ln(nC^{3}/e^{3\kappa|x|})\Big{)}\,\mathrm{d}t=\sqrt{\frac{2}{\pi}}\iint_{\mathbb{R}_{+}^{2}}e^{-\lambda t-\frac{u^{2}}{2}}1_{\{u\geq\alpha_{n}/\sqrt{t}\}}\,\mathrm{d}t\,\mathrm{d}u

where αn=13κln(nC3/e3κ|x|)\alpha_{n}=\frac{1}{3\kappa}\,\ln(nC^{3}/e^{3\kappa|x|}). Hence

v¯nε(λ)2π1λ0eλαn2u2u22du.\displaystyle\overline{v}_{n}^{\varepsilon}(\lambda)\leq\sqrt{\frac{2}{\pi}}\frac{1}{\lambda}\int_{0}^{\infty}e^{-\frac{\lambda\alpha_{n}^{2}}{u^{2}}-\frac{u^{2}}{2}}\,\mathrm{d}u.

By the change of variable u=r(2λαn2)1/4u=\sqrt{r}(2\lambda\alpha_{n}^{2})^{1/4}, we have

v¯nε(λ)\displaystyle\overline{v}_{n}^{\varepsilon}(\lambda) (2λαn2)1/4λ2π01reαnλ2(1r+r)dr\displaystyle\leq\frac{(2\lambda\alpha_{n}^{2})^{1/4}}{\lambda\sqrt{2\pi}}\int_{0}^{\infty}\frac{1}{\sqrt{r}}\,e^{-\alpha_{n}\sqrt{\frac{\lambda}{2}}(\frac{1}{r}+r)}\,\mathrm{d}r
=(2λαn2)1/4λ2π{011reαnλ2(1r+r)dr+11reαnλ2(1r+r)dr}.\displaystyle=\frac{(2\lambda\alpha_{n}^{2})^{1/4}}{\lambda\sqrt{2\pi}}\Big{\{}\int_{0}^{1}\frac{1}{\sqrt{r}}\,e^{-\alpha_{n}\sqrt{\frac{\lambda}{2}}(\frac{1}{r}+r)}\,\mathrm{d}r+\int_{1}^{\infty}\frac{1}{\sqrt{r}}\,e^{-\alpha_{n}\sqrt{\frac{\lambda}{2}}(\frac{1}{r}+r)}\,\mathrm{d}r\Big{\}}.

Using the change of variable r1rr\mapsto\frac{1}{r} in the first integral leads to

v¯nε(λ)\displaystyle\overline{v}_{n}^{\varepsilon}(\lambda) (2λαn2)1/4λ2π1(1r+1r3/2)eαnλ2(1r+r)dr2(2λαn2)1/4λ2π1eαnλ2rdr\displaystyle\leq\frac{(2\lambda\alpha_{n}^{2})^{1/4}}{\lambda\sqrt{2\pi}}\int_{1}^{\infty}\Big{(}\frac{1}{\sqrt{r}}+\frac{1}{r^{3/2}}\Big{)}\,e^{-\alpha_{n}\sqrt{\frac{\lambda}{2}}(\frac{1}{r}+r)}\,\mathrm{d}r\leq\frac{2(2\lambda\alpha_{n}^{2})^{1/4}}{\lambda\sqrt{2\pi}}\int_{1}^{\infty}\,e^{-\alpha_{n}\sqrt{\frac{\lambda}{2}}r}\,\mathrm{d}r
1παn(2λ)5/4eαnλ2.\displaystyle\leq\frac{1}{\sqrt{\pi\alpha_{n}}}\,\Big{(}\frac{2}{\lambda}\Big{)}^{5/4}\,e^{-\alpha_{n}\sqrt{\frac{\lambda}{2}}}.

Since αn13κlnn\alpha_{n}\sim\frac{1}{3\kappa}\,\ln n, the upper bound is a term of a convergent series as soon as λ>λ0:=18κ2\lambda>\lambda_{0}:=18\kappa^{2}. Therefore, by comparison,

n1v¯nε(λ)<forλ>λ0.\sum_{n\geq 1}\overline{v}_{n}^{\varepsilon}(\lambda)<\infty\quad\mbox{for}\quad\lambda>\lambda_{0}. (3.20)

Combining (3.14), (3.17) and (3.19) leads to the announced statement ψε(t,x)<\psi^{\varepsilon}(t,x)<\infty. Since

0eλtψε(t,x)dt=n1u¯nε(λ)+n1v¯nε(λ),\int_{0}^{\infty}e^{-\lambda t}\psi^{\varepsilon}(t,x)\,\mathrm{d}t=\sum_{n\geq 1}\overline{u}_{n}^{\varepsilon}(\lambda)+\sum_{n\geq 1}\overline{v}_{n}^{\varepsilon}(\lambda),

the convergence (3.18) and (3.20) of both series for λ>λ0\lambda>\lambda_{0} implies that the Laplace transform is well defined for λ>λ0\lambda>\lambda_{0}. Of course, this result can be extended for complex values λ\lambda\in\mathbb{C} satisfying Re(λ)>λ0{\rm Re}(\lambda)>\lambda_{0}. ∎

Proposition 3.10. —

Under Assumption 3.6, the function (t,x)ψε(t,x)=𝔼[Ntε|x0ε=x](t,x)\mapsto\psi^{\varepsilon}(t,x)=\mathbb{E}[N^{\varepsilon}_{t}|x_{0}^{\varepsilon}=x] is continuous.

Proof.

Let us consider the Brownian skeleton (BS)η({\rm BS})_{\eta} which corresponds to the sequences (Unε)n1(U_{n}^{\varepsilon})_{n\geq 1}, (snε)n1(s_{n}^{\varepsilon})_{n\geq 1} and (xnε)n0(x_{n}^{\varepsilon})_{n\geq 0} with x0ε=xx_{0}^{\varepsilon}=x. We consider also a second Brownian approximation (U^nε)n1(\hat{U}_{n}^{\varepsilon})_{n\geq 1}, (s^nε)n1(\hat{s}_{n}^{\varepsilon})_{n\geq 1} and (x^nε)n0(\hat{x}_{n}^{\varepsilon})_{n\geq 0} with x^0ε=x^\hat{x}_{0}^{\varepsilon}=\hat{x}, both approximations being constructed with respect to the same r.v. (An)n1(A_{n})_{n\geq 1}, (Zn)n1(Z_{n})_{n\geq 1} and the same function η\eta. The corresponding counting processes are denoted by NtεN_{t}^{\varepsilon} and N^tε\hat{N}_{t}^{\varepsilon}.
Step 1. Let us describe the distance between these two schemes. The function η\eta is bounded so we denote by M=supxη(x)M=\sup_{x\in\mathbb{R}}\eta(x) and LLipL_{\rm Lip} the Lipschitz constant of η\eta. Hence

|η2(xnε)η2(x^nε)|2MLLip|xnεx^nε|,n0.|\eta^{2}(x_{n}^{\varepsilon})-\eta^{2}(\hat{x}_{n}^{\varepsilon})|\leq 2ML_{\rm Lip}|x_{n}^{\varepsilon}-\hat{x}_{n}^{\varepsilon}|,\quad\forall n\geq 0.

Using the definition of the approximations (BS)η({\rm BS})_{\eta}, we have

|xnεx^nε|\displaystyle|x_{n}^{\varepsilon}-\hat{x}_{n}^{\varepsilon}| =|xn1εx^n1ε+εZnϕ1(e1An)(η(xn1ε)η(x^n1ε))|\displaystyle=|x_{n-1}^{\varepsilon}-\hat{x}_{n-1}^{\varepsilon}+\varepsilon Z_{n}\phi_{1}(e^{1-A_{n}})(\eta(x_{n-1}^{\varepsilon})-\eta(\hat{x}_{n-1}^{\varepsilon}))|
|xn1εx^n1ε|+ε|η(xn1ε)η(xn1ε)|\displaystyle\leq|x_{n-1}^{\varepsilon}-\hat{x}_{n-1}^{\varepsilon}|+\varepsilon|\eta(x_{n-1}^{\varepsilon})-\eta(x_{n-1}^{\varepsilon})|
(1+εLLip)|xn1εx^n1ε|(1+εLLip)n|x0εx^0ε|.\displaystyle\leq(1+\varepsilon L_{\rm Lip})|x_{n-1}^{\varepsilon}-\hat{x}_{n-1}^{\varepsilon}|\leq(1+\varepsilon L_{\rm Lip})^{n}|x_{0}^{\varepsilon}-\hat{x}_{0}^{\varepsilon}|.

We deduce

max0kn|η2(xkε)η2(x^kε)|2MLLip(1+εLLip)n|xx^|.\max_{0\leq k\leq n}|\eta^{2}(x_{k}^{\varepsilon})-\eta^{2}(\hat{x}_{k}^{\varepsilon})|\leq 2ML_{\rm Lip}(1+\varepsilon L_{\rm Lip})^{n}|x-\hat{x}|. (3.21)

Step 2. Since NtεN_{t}^{\varepsilon} is a \mathbb{N}-valued random variable, we get

ψε(t,x)\displaystyle\psi^{\varepsilon}(t,x) =n1(Ntεn)=n1(snεt)=n1(snεt,s^nεt)+n1(snεt,s^nε>t)\displaystyle=\sum_{n\geq 1}\mathbb{P}(N_{t}^{\varepsilon}\geq n)=\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t)=\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,\hat{s}_{n}^{\varepsilon}\leq t)+\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,\hat{s}_{n}^{\varepsilon}>t)
=n1(s^nεt)n1(s^nεt,snε>t)+n1(snεt,s^nε>t).\displaystyle=\sum_{n\geq 1}\mathbb{P}(\hat{s}_{n}^{\varepsilon}\leq t)-\sum_{n\geq 1}\mathbb{P}(\hat{s}_{n}^{\varepsilon}\leq t,\,s_{n}^{\varepsilon}>t)+\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,\hat{s}_{n}^{\varepsilon}>t).

We deduce that

|ψε(t,x)ψε(t,x^)|n1unε(t)+n1u¯nε(t),|\psi^{\varepsilon}(t,x)-\psi^{\varepsilon}(t,\hat{x})|\leq\sum_{n\geq 1}u_{n}^{\varepsilon}(t)+\sum_{n\geq 1}\overline{u}_{n}^{\varepsilon}(t), (3.22)

where unε(t)=(s^nεt,snε>t)u_{n}^{\varepsilon}(t)=\mathbb{P}(\hat{s}_{n}^{\varepsilon}\leq t,\,s_{n}^{\varepsilon}>t) and u¯nε(t)=(snεt,s^nε>t)\overline{u}_{n}^{\varepsilon}(t)=\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,\hat{s}_{n}^{\varepsilon}>t). Since u¯nε(t)(snεt)\overline{u}_{n}^{\varepsilon}(t)\leq\mathbb{P}(s_{n}^{\varepsilon}\leq t) which is the term of a convergent series (see Proposition 3.9), then for any ρ>0\rho>0 there exists n0εn_{0}^{\varepsilon}\in\mathbb{N} such that

n>n0εu¯nε(t)<ρ.\sum_{n>n_{0}^{\varepsilon}}\overline{u}_{n}^{\varepsilon}(t)<\rho. (3.23)

Moreover

u¯nε(t)=(snεt,snε+(s^nεsnε)>t)(tδ<snεt)+(s^nεsnε>δ).\overline{u}_{n}^{\varepsilon}(t)=\mathbb{P}(s_{n}^{\varepsilon}\leq t,\,s_{n}^{\varepsilon}+(\hat{s}_{n}^{\varepsilon}-s_{n}^{\varepsilon})>t)\leq\mathbb{P}(t-\delta<s_{n}^{\varepsilon}\leq t)+\mathbb{P}(\hat{s}_{n}^{\varepsilon}-s_{n}^{\varepsilon}>\delta).

The random variables (snε)1nn0ε(s_{n}^{\varepsilon})_{1\leq n\leq n_{0}^{\varepsilon}} are absolutely continuous with respect to the Lebesgue measure. Consequently

ρ>0,δε>0such thatn=1n0ε(tδε<snεt)ρ.\forall\rho>0,\quad\exists\delta_{\varepsilon}>0\quad\mbox{such that}\quad\sum_{n=1}^{n_{0}^{\varepsilon}}\mathbb{P}(t-\delta_{\varepsilon}<s_{n}^{\varepsilon}\leq t)\leq\rho. (3.24)

It suffices therefore to deal with the remaining expression: n=1n0ε(s^nεsnε>δε)\sum_{n=1}^{n_{0}^{\varepsilon}}\mathbb{P}(\hat{s}_{n}^{\varepsilon}-s_{n}^{\varepsilon}>\delta_{\varepsilon}). By Step 1 of the proof and by the definition of (snε)(s_{n}^{\varepsilon}) and (s^nε)(\hat{s}_{n}^{\varepsilon}), we have for nn0εn\leq n_{0}^{\varepsilon}

|snεs^nε|\displaystyle|s_{n}^{\varepsilon}-\hat{s}_{n}^{\varepsilon}| =ε2|(η2(x0ε)η2(x^0ε))e1A1++(η2(xn1ε)η2(x^n1ε))e1An|\displaystyle=\varepsilon^{2}|(\eta^{2}(x_{0}^{\varepsilon})-\eta^{2}(\hat{x}_{0}^{\varepsilon}))e^{1-A_{1}}+\ldots+(\eta^{2}(x_{n-1}^{\varepsilon})-\eta^{2}(\hat{x}_{n-1}^{\varepsilon}))e^{1-A_{n}}|
ε2max0kn1|η2(xkε)η2(x^kε))|(e1A1++e1An)\displaystyle\leq\varepsilon^{2}\,\max_{0\leq k\leq n-1}|\eta^{2}(x_{k}^{\varepsilon})-\eta^{2}(\hat{x}_{k}^{\varepsilon}))|(e^{1-A_{1}}+\ldots+e^{1-A_{n}})
Cε2|xx^|(e1A1++e1An),\displaystyle\leq C\,\varepsilon^{2}|x-\hat{x}|(e^{1-A_{1}}+\ldots+e^{1-A_{n}}),

with C=2MLLip(1+εLLip)n0εC=2ML_{\rm Lip}(1+\varepsilon L_{\rm Lip})^{n_{0}^{\varepsilon}}. Hence

ρ>0,κε>0such that|xx^|<κεn=1n0ε(s^nεsnε>δε)ρ.\forall\rho>0,\quad\exists\kappa_{\varepsilon}>0\quad\mbox{such that}\quad|x-\hat{x}|<\kappa_{\varepsilon}\ \Rightarrow\ \sum_{n=1}^{n_{0}^{\varepsilon}}\mathbb{P}(\hat{s}_{n}^{\varepsilon}-s_{n}^{\varepsilon}>\delta_{\varepsilon})\leq\rho. (3.25)

Combining (3.23), (3.24) and (3.25), we obtain

ρ>0,κε>0such that|xx^|<κεn1u¯nε(t)3ρ.\forall\rho>0,\quad\exists\kappa_{\varepsilon}>0\quad\mbox{such that}\quad|x-\hat{x}|<\kappa_{\varepsilon}\ \Rightarrow\ \sum_{n\geq 1}\overline{u}_{n}^{\varepsilon}(t)\leq 3\rho.

Let us use now similar arguments in order to bound the series associated to unε(t)u_{n}^{\varepsilon}(t). Using the following upper-bound,

unε(t)=(snε>t,snε+(s^nεsnε)t)(t<snεt+δ)+(s^nεsnε>δ),u_{n}^{\varepsilon}(t)=\mathbb{P}(s_{n}^{\varepsilon}>t,\,s_{n}^{\varepsilon}+(\hat{s}_{n}^{\varepsilon}-s_{n}^{\varepsilon})\leq t)\leq\mathbb{P}(t<s_{n}^{\varepsilon}\leq t+\delta)+\mathbb{P}(\hat{s}_{n}^{\varepsilon}-s_{n}^{\varepsilon}>\delta),

we deduce that all arguments presented so far and concerning u¯nε(t)\overline{u}_{n}^{\varepsilon}(t) can be used for unε(t)u_{n}^{\varepsilon}(t). Finally (3.22) leads to the continuity of xψε(t,x)x\mapsto\psi^{\varepsilon}(t,x):

ρ>0,κε>0such that|xx^|<κε|ψε(t,x)ψε(t,x^)|6ρ.\forall\rho>0,\quad\exists\kappa_{\varepsilon}>0\quad\mbox{such that}\quad|x-\hat{x}|<\kappa_{\varepsilon}\ \Rightarrow\ |\psi^{\varepsilon}(t,x)-\psi^{\varepsilon}(t,\hat{x})|\leq 6\rho.

Let us end the proof by focusing our attention on the continuity with respect to the time variable. Since

ψε(t,x)=n1(snεt),\psi^{\varepsilon}(t,x)=\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t),

where snεs_{n}^{\varepsilon} is an absolutely continuous random variable, the Lebesgue monotone convergence theorem implies the continuity of tψε(t,x)t\mapsto\psi^{\varepsilon}(t,x). ∎

We give now an important result concerning the function ψε(t,x)=𝔼[Ntε|x0ε=x]\psi^{\varepsilon}(t,x)=\mathbb{E}[N_{t}^{\varepsilon}|x_{0}^{\varepsilon}=x].

Proposition 3.11. —

Under Assumption 3.5 and Assumption 3.6 , there exist CT>0C_{T}>0, κ>0\kappa>0 and ε0>0\varepsilon_{0}>0 such that

ε2ψε(t,x)CTeκ|x|,t[0,T],x,εε0.\varepsilon^{2}\psi^{\varepsilon}(t,x)\leq C_{T}\,e^{\kappa|x|},\quad\forall t\in[0,T],\ \forall x\in\mathbb{R},\ \forall\varepsilon\leq\varepsilon_{0}. (3.26)

Let us note that the constant κ\kappa is explicit: it suffices to choose κ=32κ2\kappa=3\sqrt{2}\kappa_{2} where κ2\kappa_{2} corresponds to the constant introduced in Assumption 3.5.

Proof.

The proof follows similar ideas as those developed in Proposition 3.9. We introduce here mnε:=min0kn1η2(xkε)m_{n}^{\varepsilon}:=\min_{0\leq k\leq n-1}\eta^{2}(x_{k}^{\varepsilon}) and recall that

ψε(t,x)=n1(snεt).\psi^{\varepsilon}(t,x)=\displaystyle\sum_{n\geq 1}\mathbb{P}(s_{n}^{\varepsilon}\leq t). (3.27)

We aim to control

ε2ψε(t,x)=n1(ε2unε(t,x)+ε2vnε(t,x))\varepsilon^{2}\psi^{\varepsilon}(t,x)=\sum_{n\geq 1}(\varepsilon^{2}u_{n}^{\varepsilon}(t,x)+\varepsilon^{2}v_{n}^{\varepsilon}(t,x)) (3.28)

by using similar notations as those presented in Proposition 3.9, that is

unε(t,x)=(snεt;mnεεγn2/3|x0ε=x),\displaystyle u_{n}^{\varepsilon}(t,x)=\mathbb{P}(s_{n}^{\varepsilon}\leq t;m_{n}^{\varepsilon}\geq\varepsilon^{\gamma}n^{-2/3}|x_{0}^{\varepsilon}=x), (3.29)
vnε(t,x)=(snεt;mnε<εγn2/3|x0ε=x).\displaystyle v_{n}^{\varepsilon}(t,x)=\mathbb{P}(s_{n}^{\varepsilon}\leq t;m_{n}^{\varepsilon}<\varepsilon^{\gamma}n^{-2/3}|x_{0}^{\varepsilon}=x). (3.30)

Here a suitable choice of the exponent γ\gamma should ensure the required boundedness. We shall discuss about this choice in the following.
Step 1. Consider first the sequence (unε(t,x))n1(u_{n}^{\varepsilon}(t,x))_{n\geq 1}.

Let us introduce nεT:=39/2(2T/e)3ε2n_{\varepsilon}^{T}:=3^{9/2}(2T/e)^{3}\varepsilon^{-2} and decompose the series associated to unεu_{n}^{\varepsilon} as follows:

n1ε2unε(t,x)=ε2n=1nεTunε(t,x)+ε2n>nεTunε(t,x)ε2nεT+ε2n>nεTunε(t,x)39/2(2T/e)3+ε2n>nεTunε(t,x).\begin{array}[]{ll}\displaystyle\sum_{n\geq 1}\varepsilon^{2}u_{n}^{\varepsilon}(t,x)&=\varepsilon^{2}\sum_{n=1}^{n_{\varepsilon}^{T}}u_{n}^{\varepsilon}(t,x)+\varepsilon^{2}\sum_{n>n_{\varepsilon}^{T}}u_{n}^{\varepsilon}(t,x)\\ \displaystyle&\displaystyle\leq\varepsilon^{2}\cdot n_{\varepsilon}^{T}+\varepsilon^{2}\sum_{n>n_{\varepsilon}^{T}}u_{n}^{\varepsilon}(t,x)\leq 3^{9/2}(2T/e)^{3}+\varepsilon^{2}\sum_{n>n_{\varepsilon}^{T}}u_{n}^{\varepsilon}(t,x).\end{array} (3.31)

We focus our attention on the indices satisfying n>nεTn>n_{\varepsilon}^{T}. Let us set γ=4/3\gamma=-4/3. For this particular choice, both the definition (3.29) and the definition of nεTn_{\varepsilon}^{T} lead to

unε(t,x)(e1A1++e1Antn2/3ε(γ+2))(e1A1++e1Ann𝔼[e1A1]Tn2/3ε(γ+2)ne33/2)(e1A1++e1Ann𝔼[e1A1]Tn2/3ε(γ+2)).\begin{array}[]{ll}\displaystyle u_{n}^{\varepsilon}(t,x)&\leq\mathbb{P}\left(e^{1-A_{1}}+\ldots+e^{1-A_{n}}\leq tn^{2/3}\varepsilon^{-(\gamma+2)}\right)\\[4.0pt] \displaystyle&\displaystyle\leq\mathbb{P}\left(e^{1-A_{1}}+\ldots+e^{1-A_{n}}-n\mathbb{E}[e^{1-A_{1}}]\leq Tn^{2/3}\varepsilon^{-(\gamma+2)}-ne3^{-3/2}\right)\\[4.0pt] \displaystyle&\displaystyle\leq\mathbb{P}\left(e^{1-A_{1}}+\ldots+e^{1-A_{n}}-n\mathbb{E}[e^{1-A_{1}}]\leq-Tn^{2/3}\varepsilon^{-(\gamma+2)}\right).\end{array}

Using Hoeffding’s inequality permits to obtain the following upper-bound:

ε2n>nεTunε(t,x)ε2n>nεTexp(2T2n4/3ε2(γ+2)ne2)ε2n1exp(2T2n1/3e2ε4/3).\displaystyle\varepsilon^{2}\sum_{n>n_{\varepsilon}^{T}}u_{n}^{\varepsilon}(t,x)\leq\varepsilon^{2}\sum_{n>n_{\varepsilon}^{T}}\exp\left(-\frac{2T^{2}n^{4/3}\varepsilon^{-2(\gamma+2)}}{ne^{2}}\right)\leq\varepsilon^{2}\sum_{n\geq 1}\exp\left(-\frac{2T^{2}n^{1/3}}{e^{2}\varepsilon^{4/3}}\right). (3.32)

Combining (3.31) and (3.32), we obtain for any ε1\varepsilon\leq 1:

n1ε2unε(t,x)39/2(2T/e)3+n1exp(2T2n1/3e2)=:CT0<+.\displaystyle\sum_{n\geq 1}\varepsilon^{2}u_{n}^{\varepsilon}(t,x)\leq 3^{9/2}(2T/e)^{3}+\sum_{n\geq 1}\exp\left(-\frac{2T^{2}\,n^{1/3}}{e^{2}}\right)=:C_{T}^{0}<+\infty.

Let us just note that the upper-bound does not depend on the space variable xx.

Step 2. Let us focus now on the second part, that is, the terms ε2vnε(t,x)\varepsilon^{2}v_{n}^{\varepsilon}(t,x). Using the properties of FF (see Assumption 3.5), there exist C>0C>0 and κ>0\kappa>0 such that η(z)Ceκ|z|\eta(z)\geq Ce^{-\kappa|z|} (the value of κ\kappa here corresponds to 2κ2\sqrt{2}\kappa_{2} where κ2\kappa_{2} is the constant appearing in Assumption 3.5)

ε2vnε(t,x)ε2(st s.t. η2(Bs+x)<εγn2/3)ε2(C2exp(2κsups[0,t]|Bs+x|)<εγn2/3)ε2(exp(sups[0,t]|Bs+x|)<(C1εγ2n1/3)1/κ),\begin{array}[]{ll}\varepsilon^{2}v_{n}^{\varepsilon}(t,x)&\leq\varepsilon^{2}\mathbb{P}\left(\exists s\leq t\mbox{ s.t. }\eta^{2}(B_{s}+x)<\varepsilon^{\gamma}n^{-2/3}\right)\\ &\leq\varepsilon^{2}\mathbb{P}\Big{(}C^{2}\exp\Big{(}-2\kappa\displaystyle\sup_{s\in[0,t]}|B_{s}+x|\Big{)}<\varepsilon^{\gamma}n^{-2/3}\Big{)}\\ &\leq\varepsilon^{2}\mathbb{P}\Big{(}\exp\Big{(}-\displaystyle\sup_{s\in[0,t]}|B_{s}+x|\Big{)}<\Big{(}C^{-1}\varepsilon^{\frac{\gamma}{2}}n^{-1/3}\Big{)}^{1/\kappa}\Big{)},\end{array} (3.33)

for all (t,x)[0,T]×(t,x)\in[0,T]\times\mathbb{R}. We need here an auxiliary result.

Lemma 3.12. —

Let us define the function

t(x,z):=(sups[0,t]|x+Bs|>ln(z)).\mathcal{R}_{t}(x,z):=\mathbb{P}\Big{(}\displaystyle\sup_{s\in[0,t]}|x+B_{s}|>\ln(z)\Big{)}. (3.34)

For any (x,z)×(0,)(x,z)\in\mathbb{R}\times(0,\infty) we have t(x,z)0\mathcal{R}_{t}(x,z)\geq 0 and zt(x,z)z\mapsto\mathcal{R}_{t}(x,z) is non increasing. Furthermore for any δ>0\delta>0, there exists CT>0C_{T}>0 such that

0+t(x,z1/δ)dzCTeδ|x|,(t,x)[0,T]×.\displaystyle\int_{0}^{+\infty}\mathcal{R}_{t}(x,z^{1/\delta})\,\mathrm{d}z\leq C_{T}\ e^{\delta|x|},\quad\forall(t,x)\in[0,T]\times\mathbb{R}. (3.35)

We postpone the proof of this lemma. We observe therefore

vnε(t,x)t(x,(Cn1/3εγ/2)1/κ)=(sups[0,t]|x+Bs|>1κln(Cn1/3εγ/2)).v_{n}^{\varepsilon}(t,x)\leq\mathcal{R}_{t}\left(x,\Big{(}\frac{Cn^{1/3}}{\varepsilon^{\gamma/2}}\Big{)}^{1/\kappa}\right)=\mathbb{P}\left(\displaystyle\sup_{s\in[0,t]}|x+B_{s}|>\frac{1}{\kappa}\ln\left(\frac{Cn^{1/3}}{\varepsilon^{\gamma/2}}\right)\right). (3.36)

We define n(ε)=inf{n0 s.t. Cn1/3εγ/2}n(\varepsilon)=\inf\{n\geq 0\mbox{ s.t. }Cn^{1/3}\geq\varepsilon^{\gamma/2}\}. Then

ε2n1vnε(t,x)=ε2n=1n(ε)vnε(t,x)+ε2nn(ε)+1vnε(t,x)ε2n(ε)+ε2nn(ε)+1t(x,(Cn1/3εγ/2)1/κ)ε2n(ε)+ε2n(ε)+t(x,(Cz1/3εγ/2)1/κ)dz.\begin{array}[]{ll}\varepsilon^{2}\displaystyle\sum_{n\geq 1}v_{n}^{\varepsilon}(t,x)&=\varepsilon^{2}\displaystyle\sum_{n=1}^{n(\varepsilon)}v_{n}^{\varepsilon}(t,x)+\varepsilon^{2}\displaystyle\sum_{n\geq n(\varepsilon)+1}v_{n}^{\varepsilon}(t,x)\\ &\leq\varepsilon^{2}n(\varepsilon)+\varepsilon^{2}\displaystyle\sum_{n\geq n(\varepsilon)+1}\mathcal{R}_{t}\left(x,\Big{(}\displaystyle\frac{Cn^{1/3}}{\varepsilon^{\gamma/2}}\Big{)}^{1/\kappa}\right)\\ &\leq\varepsilon^{2}n(\varepsilon)+\varepsilon^{2}\displaystyle\int_{n(\varepsilon)}^{+\infty}\mathcal{R}_{t}\left(x,\Big{(}\displaystyle\frac{Cz^{1/3}}{\varepsilon^{\gamma/2}}\Big{)}^{1/\kappa}\right)\mathrm{d}z.\end{array} (3.37)

In the last expression only the term under the integral depends on xx. We perform the change of variable in this term of the form y=ε2zy=\varepsilon^{2}z and obtain:

ε2n1vnε(t,x)ε2n(ε)+ε2n(ε)+t(x,(Cy1/3ε2/3+γ/2)1/κ)dyε2n(ε)+0+t(x,C1/κy1/3κ)dyε2n(ε)+CTe3κ|x|,\begin{array}[]{ll}\varepsilon^{2}\displaystyle\sum_{n\geq 1}v_{n}^{\varepsilon}(t,x)&\leq\varepsilon^{2}n(\varepsilon)+\displaystyle\int_{\varepsilon^{2}n(\varepsilon)}^{+\infty}\mathcal{R}_{t}\left(x,\Big{(}\displaystyle\frac{Cy^{1/3}}{\varepsilon^{2/3+\gamma/2}}\Big{)}^{1/\kappa}\right)\mathrm{d}y\\ &\leq\varepsilon^{2}n(\varepsilon)+\displaystyle\int_{0}^{+\infty}\mathcal{R}_{t}(x,C^{1/\kappa}y^{1/3\kappa})\mathrm{d}y\leq\varepsilon^{2}n(\varepsilon)+C_{T}e^{3\kappa|x|},\end{array} (3.38)

by using the particular value γ=4/3\gamma=-4/3 and Lemma 3.12. In order to conclude we need to control ε2n(ε)\varepsilon^{2}n(\varepsilon). By the definition of n(ε)n(\varepsilon) we have

n(ε)=inf{n0 s.t. n1ε2C3}1ε2C3+1,forε1.n(\varepsilon)=\displaystyle\inf\left\{n\geq 0\mbox{ s.t. }n\geq\displaystyle\frac{1}{\varepsilon^{2}C^{3}}\right\}\leq\frac{1}{\varepsilon^{2}C^{3}}+1,\quad\mbox{for}\ \varepsilon\leq 1. (3.39)

This allows us to conclude that ε2n(ε)1C3+1\varepsilon^{2}n(\varepsilon)\leq\frac{1}{C^{3}}+1 for ε1\varepsilon\leq 1. Combining the two steps of the proof leads to the announced upper-bound (3.26). ∎

Proof of Lemma 3.12.

The proof of the first two properties is obvious by using the definition of t\mathcal{R}_{t}. Let us show that (3.35) is true. By using the reflection principle of the Brownian motion we can evaluate

0+t(x,z1/δ)dz=0eδ(1+|x|)t(x,z1/δ)dz+eδ(1+|x|)+t(x,z1/δ)dzeδ(1+|x|)+4eδ(1+|x|)+(G>1t(1δln(z)|x|))dzeδ(1+|x|)+4eδ(1+|x|)+(G>1T(1δln(z)|x|))dz,\begin{array}[]{ll}\displaystyle\int_{0}^{+\infty}\mathcal{R}_{t}(x,z^{1/\delta})\mathrm{d}z&=\displaystyle\int_{0}^{e^{\delta(1+|x|)}}\mathcal{R}_{t}(x,z^{1/\delta})\mathrm{d}z+\displaystyle\int_{e^{\delta(1+|x|)}}^{+\infty}\mathcal{R}_{t}(x,z^{1/\delta})\mathrm{d}z\\ &\leq e^{\delta(1+|x|)}+4\displaystyle\int_{e^{\delta(1+|x|)}}^{+\infty}\mathbb{P}\Big{(}G>\displaystyle\frac{1}{\sqrt{t}}\Big{(}\displaystyle\frac{1}{\delta}\ln(z)-|x|\Big{)}\Big{)}\mathrm{d}z\\ &\leq e^{\delta(1+|x|)}+4\displaystyle\int_{e^{\delta(1+|x|)}}^{+\infty}\mathbb{P}\Big{(}G>\displaystyle\frac{1}{\sqrt{T}}\Big{(}\displaystyle\frac{1}{\delta}\ln(z)-|x|\Big{)}\Big{)}\mathrm{d}z,\end{array} (3.40)

where GG denotes a standard normal random variable 𝒩(0,1){\cal{N}}(0,1). We used the fact that for any u>0u>0 we have (Gu)1u2πeu22\mathbb{P}(G\geq u)\leq\frac{1}{u\sqrt{2\pi}}e^{-\frac{u^{2}}{2}}. Hence, for zeδ(1+|x|)z\geq e^{\delta(1+|x|)},

(G>1T(1δln(z)|x|))\displaystyle\mathbb{P}\Big{(}G>\displaystyle\frac{1}{\sqrt{T}}\Big{(}\displaystyle\frac{1}{\delta}\ln(z)-|x|\Big{)}\Big{)} δT(ln(z)δ|x|)2πexp{12T(1δln(z)|x|)2}\displaystyle\leq\frac{\delta\sqrt{T}}{(\ln(z)-\delta|x|)\sqrt{2\pi}}\,\exp\Big{\{}-\frac{1}{2T}\Big{(}\frac{1}{\delta}\,\ln(z)-|x|\Big{)}^{2}\Big{\}}
T2πexp{12T(1δln(z)|x|)2}.\displaystyle\leq\frac{\sqrt{T}}{\sqrt{2\pi}}\,\exp\Big{\{}-\frac{1}{2T}\Big{(}\frac{1}{\delta}\,\ln(z)-|x|\Big{)}^{2}\Big{\}}.

We deduce

0+t(x,z1/δ)dz\displaystyle\displaystyle\int_{0}^{+\infty}\mathcal{R}_{t}(x,z^{1/\delta})\mathrm{d}z eδ(1+|x|)+4T2πeδ(1+|x|)+exp{12T(1δln(z)|x|)2}dz.\displaystyle\leq e^{\delta(1+|x|)}+\frac{4\sqrt{T}}{\sqrt{2\pi}}\displaystyle\int_{e^{\delta(1+|x|)}}^{+\infty}\,\exp\Big{\{}-\frac{1}{2T}\Big{(}\frac{1}{\delta}\,\ln(z)-|x|\Big{)}^{2}\Big{\}}\mathrm{d}z.

By doing the change of variable z=eδ(1+|x|)yz=e^{\delta(1+|x|)}y, we have

0+t(x,z1/δ)dz\displaystyle\displaystyle\int_{0}^{+\infty}\mathcal{R}_{t}(x,z^{1/\delta})\mathrm{d}z eδ(1+|x|)[1+4T2π1+e12T(1δ(δ(1+|x|)+lny)|x|)2dy]\displaystyle\leq e^{\delta(1+|x|)}\Big{[}1+\displaystyle\frac{4\sqrt{T}}{\sqrt{2\pi}}\displaystyle\int_{1}^{+\infty}e^{\displaystyle-\frac{1}{2T}(\frac{1}{\delta}(\delta(1+|x|)+\ln y)-|x|)^{2}}\mathrm{d}y\Big{]}
eδ(1+|x|)[1+4T2π1+e12T(1+1δln(y))2dy]=:CTeδ|x|.\displaystyle\leq e^{\delta(1+|x|)}\Big{[}1+\displaystyle\frac{4\sqrt{T}}{\sqrt{2\pi}}\displaystyle\int_{1}^{+\infty}e^{-\displaystyle\frac{1}{2T}(1+\frac{1}{\delta}\ln(y))^{2}}\mathrm{d}y\Big{]}=:C_{T}\,e^{\delta|x|}.

The upper-bound holds for any (t,x)[0,T]×(t,x)\in[0,T]\times\mathbb{R} as announced. ∎

Since Proposition 3.9 and Proposition 3.10 point out different preliminary properties of the average number of steps needed by the Brownian skeleton to cover the time interval [0,T][0,T], the study of the ε\varepsilon-strong approximation of both the linear and growth diffusions can be achieved.

Theorem 3.13. —

Let (Xt)0tT(X_{t})_{0\leq t\leq T} be a solution of the stochastic differential equation (3.1) satisfying both Assumptions 3.5 and 3.6. Let (ytε)0tT(y^{\varepsilon}_{t})_{0\leq t\leq T} be the ε\varepsilon-strong approximation of (Xt)0tT(X_{t})_{0\leq t\leq T} given by (3.12) and N^Tε\hat{N}^{\varepsilon}_{T} the random number of points needed to build this approximation. Then, there exist μ>0\mu>0 such that

limε0ε2𝔼[N^Tε]=μ𝔼[0ρ(T)1η2(x+Bs)ds],(T,x)+×\lim_{\varepsilon\to 0}\varepsilon^{2}\,\mathbb{E}[\hat{N}^{\varepsilon}_{T}]=\mu\,\mathbb{E}\left[\displaystyle\int_{0}^{\rho(T)}\frac{1}{\eta^{2}(x+B_{s})}\mathrm{d}s\right],\quad\forall(T,x)\in\mathbb{R}_{+}\times\mathbb{R} (3.41)

where (Bt)t0(B_{t})_{t\geq 0} stands for a standard one-dimensional Brownian motion.

Remark 3.14. —
  1. 1.

    The constant appearing in the statement is explicitly known. Let us introduce MM the cumulative distribution function associated to the random variable e1Ae^{1-A} with AΓ(3/2,2)A\sim{\Gamma}(3/2,2). We denote by M(f)=0f(s)dM(s)M(f)=\int_{0}^{\infty}f(s)\,\mathrm{d}M(s), for any nonnegative function ff. Then

    μ=1M(ϕ12)andM(ϕ12)=M(Id)=e 33/20.5231336.\mu=\frac{1}{M(\phi_{1}^{2})}\quad\mbox{and}\quad M(\phi_{1}^{2})=M({\rm Id})=e\,3^{-3/2}\approx 0.5231336. (3.42)
  2. 2.

    Let us note that the link between the function η\eta, defining the approximation scheme, and the function FF introduced in Assumption 3.5, permits to write

    limε0ε2𝔼[N^Tε]\displaystyle\lim_{\varepsilon\to 0}\varepsilon^{2}\,\mathbb{E}[\hat{N}^{\varepsilon}_{T}] =(eκmin1+1)2μ𝔼[0ρ(T)F2(2(x+Bs)2+1)ds]\displaystyle=(e\kappa_{\rm min}^{-1}+1)^{2}\,\mu\,\mathbb{E}\left[\displaystyle\int_{0}^{\rho(T)}F^{2}(2(x+B_{s})^{2}+1)\mathrm{d}s\right]
    =(eκmin1+1)2μ𝔼[F2(2(x+y)2+1)Lρ(T)(y)dy].\displaystyle=(e\kappa_{\rm min}^{-1}+1)^{2}\,\mu\,\mathbb{E}\left[\displaystyle\int_{\mathbb{R}}F^{2}(2(x+y)^{2}+1)L_{\rho(T)}(y)\mathrm{d}y\right].

    The last equality is just an immediate application of the occupation time formula (see, for instance, Corollary 1.6 page 209 in [21]), Lt(y)L_{t}(y) standing for the local time of the standard Brownian motion. A proof of Theorem 3.13 based on the local times of the Brownian motion and therefore on a precise description of the Brownian paths could be investigated, we prefer here to propose a proof involving a renewal property of the average number of points in the numerical scheme.

Proof of Theorem 3.13.

We start to mention that the notation of the constants is generic through this proof: C:=CθC:=C_{\theta} or κ:=κθ\kappa:=\kappa_{\theta} if the constants depend on a parameter θ\theta.
The proof of the theorem is based on the study of a particular renewal inequality. The material is organized as follows: on one hand we shall prove that uε(t,x):=ε2𝔼[Ntε|x0ε=x]u^{\varepsilon}(t,x):=\varepsilon^{2}\mathbb{E}[N_{t}^{\varepsilon}|x_{0}^{\varepsilon}=x], NtεN^{\varepsilon}_{t} being defined in (3.13), satisfies a renewal equation. On the other hand, we describe U(t,x)U(t,x) defined by

U(t,x):=μ𝔼[0t1η2(x+Bs)ds]=(eκmin1+1)2M(ϕ12)𝔼[0tF2(2(x+Bs)2+1)ds],U(t,x):=\mu\mathbb{E}\left[\displaystyle\int_{0}^{t}\frac{1}{\eta^{2}(x+B_{s})}\mathrm{d}s\right]=\frac{(e\kappa_{\rm min}^{-1}+1)^{2}}{M(\phi_{1}^{2})}\,\mathbb{E}\left[\displaystyle\int_{0}^{t}F^{2}(2(x+B_{s})^{2}+1)\mathrm{d}s\right], (3.43)

where μ\mu corresponds to the constant introduced in the statement of the theorem and described in Remark 3.14. Then we observe that the difference:

Dε(t,x):=uε(t,x)U(t,x)D^{\varepsilon}(t,x):=u^{\varepsilon}(t,x)-U(t,x) (3.44)

satisfies a renewal inequality which leads to limε0Dε(t,x)=0\lim_{\varepsilon\to 0}D^{\varepsilon}(t,x)=0.
Step 1. Renewal equation satisfied by ψε(t,x)=𝔼[Ntε|x𝟎ε=x]\psi^{\varepsilon}(t,x)=\mathbb{E}[N^{\varepsilon}_{t}|x_{0}^{\varepsilon}=x]. Let us note that ψε(t,x)\psi^{\varepsilon}(t,x) satisfies the following renewal equation:

ψε(t,x)=M(t/(ε2η2(x)))+i=±1120t/(ε2η2(x))ψε(tsε2η2(x),x+iεη(x)ϕ1(s))dM(s)\psi^{\varepsilon}(t,x)=M(t/(\varepsilon^{2}\eta^{2}(x)))+\sum_{i=\pm 1}\frac{1}{2}\int_{0}^{t/(\varepsilon^{2}\eta^{2}(x))}\psi^{\varepsilon}\Big{(}t-s\varepsilon^{2}\eta^{2}(x),\ x+i\varepsilon\eta(x)\phi_{1}(s)\Big{)}\,\mathrm{d}M(s) (3.45)

where MM corresponds to the cumulative distribution function associated to the random variable e1Ae^{1-A} with AΓ(3/2,2)A\sim{\Gamma}(3/2,2). Indeed, we focus our attention to U1εU_{1}^{\varepsilon} the first positive abscissa of the Brownian paths skeleton (BS)η({\rm BS})_{\eta}.

We can observe two possibilities:

  • Either U1ε>tU_{1}^{\varepsilon}>t and consequently Ntε=0N_{t}^{\varepsilon}=0.

  • Either U1ε=sε2η2(x)tU_{1}^{\varepsilon}=s\varepsilon^{2}\eta^{2}(x)\leq t. The Markov property implies the following identity in distribution: for any non negative measurable function \mathcal{H}, we have

    𝔼[(Ntε)|x0ε,U1ε]=𝔼[0(tsε2η2(x),x1ε)|x0ε,U1ε]with0(t,y)=𝔼[(1+Ntε)|x0ε=y].\mathbb{E}[\mathcal{H}(N_{t}^{\varepsilon})|x_{0}^{\varepsilon},\,U_{1}^{\varepsilon}]=\mathbb{E}[\mathcal{H}_{0}(t-s\varepsilon^{2}\eta^{2}(x),x_{1}^{\varepsilon})|x_{0}^{\varepsilon},\,U_{1}^{\varepsilon}]\ \mbox{with}\ \mathcal{H}_{0}(t,y)=\mathbb{E}[\mathcal{H}(1+N^{\varepsilon}_{t})|x_{0}^{\varepsilon}=y].

    We just note that the function 0(t,y)\mathcal{H}_{0}(t,y) associated with (y)=y1y0\mathcal{H}(y)=y1_{y\geq 0} corresponds to 1+ψε(t,y)1+\psi^{\varepsilon}(t,y).

Hence

ψε(t,x)\displaystyle\psi^{\varepsilon}(t,x) =M(t/(ε2η2(x))+0t/(ε2η2(x))𝔼[ψε(tsε2η2(x),x1ε)|x0ε=x]dM(s).\displaystyle=M(t/(\varepsilon^{2}\eta^{2}(x))+\int_{0}^{t/(\varepsilon^{2}\eta^{2}(x))}\mathbb{E}\left[\psi^{\varepsilon}(t-s\varepsilon^{2}\eta^{2}(x),x_{1}^{\varepsilon})\,\Big{|}x_{0}^{\varepsilon}=x\right]\,\mathrm{d}M(s).

Let us now introduce uε(t,x):=ε2ψε(t,x)u^{\varepsilon}(t,x):=\varepsilon^{2}\psi^{\varepsilon}(t,x) and define for any nonnegative function hh:

εh(t,x)=i=±1120t/(ε2η2(x))h(tsε2η2(x),x+iεη(x)ϕ1(s))dM(s).\mathcal{M}_{\varepsilon}h(t,x)=\sum_{i=\pm 1}\frac{1}{2}\int_{0}^{t/(\varepsilon^{2}\eta^{2}(x))}h\Big{(}t-s\varepsilon^{2}\eta^{2}(x),\ x+i\varepsilon\eta(x)\phi_{1}(s)\Big{)}\,\mathrm{d}M(s). (3.46)

Then the following renewal equation holds

uε(t,x)\displaystyle u^{\varepsilon}(t,x) =ε2M(t/(ε2η2(x)))+εuε(t,x),(t,x)+×.\displaystyle=\varepsilon^{2}M(t/(\varepsilon^{2}\eta^{2}(x)))+\mathcal{M}_{\varepsilon}u^{\varepsilon}(t,x),\quad\forall(t,x)\in\mathbb{R}_{+}\times\mathbb{R}. (3.47)

Step 2. Description of the function U(t,x)U(t,x) introduced in (3.43). Due to Assumption 3.5, 𝑭F is assumed to be a 𝓒𝟐()\mathcal{C}^{2}(\mathbb{R})-continuous function and 𝑭F, 𝑭F^{\prime} and 𝑭′′F^{\prime\prime} have at most exponential growth. The dominated convergence theorem permits therefore to obtain that 𝒙𝑼(𝒕,𝒙)x\mapsto U(t,x) is also a 𝓒𝟐(,)\mathcal{C}^{2}(\mathbb{R},\mathbb{R})-continuous function. Moreover, combining Itô’s formula and Lebesgue’s theorem lead to the regularity with respect to both variables 𝒕t and 𝒙x: 𝑼U is 𝓒𝟐(+×,)\mathcal{C}^{2}(\mathbb{R}_{+}\times\mathbb{R},\mathbb{R})-continuous. Since 𝑼U is regular and has at most exponential growth, it corresponds to the probabilistic representation (see for example Karatzas and Shreve [18], p. 270, Corollary 4.5) of the unique solution:

𝑴(𝐈𝐝)𝑼𝒕(𝒕,𝒙)=𝑴(ϕ𝟏𝟐)𝟐𝟐𝑼𝒙𝟐(𝒕,𝒙)+(𝒆𝜿𝐦𝐢𝐧𝟏+𝟏)𝟐𝑭𝟐(𝟐𝒙𝟐+𝟏),𝑼(𝟎,𝒙)=𝟎.M({\rm Id})\frac{\partial U}{\partial t}(t,x)=\frac{M(\phi_{1}^{2})}{2}\frac{\partial^{2}U}{\partial x^{2}}(t,x)+(e\kappa_{\rm min}^{-1}+1)^{2}F^{2}(2x^{2}+1),\quad U(0,x)=0. (3.48)

We just recall that 𝑴(𝐈𝐝)=𝑴(ϕ𝟏𝟐)M({\rm Id})=M(\phi_{1}^{2}) (see Remark 3.14).
Let 𝑹>𝟎R>0. Using the Taylor expansion in order to compute the operator defined in (3.46), we obtain

𝓜𝜺𝑼(𝒕,𝒙)=\displaystyle\mathcal{M}_{\varepsilon}U(t,x)= 𝑼(𝒕,𝒙)𝑴(𝒕/(𝜺𝟐𝜼𝟐(𝒙)))𝜺𝟐𝜼𝟐(𝒙)𝑴(𝐈𝐝)𝑼𝒕(𝒕,𝒙)\displaystyle U(t,x)M(t/(\varepsilon^{2}\eta^{2}(x)))-\varepsilon^{2}\eta^{2}(x)M({\rm Id})\frac{\partial U}{\partial t}(t,x)
+𝜺𝟐𝟐𝜼𝟐(𝒙)𝑴(ϕ𝟏𝟐)𝟐𝑼𝒙𝟐(𝒕,𝒙)+𝜺𝟐𝒐𝑹(𝟏),\displaystyle+\frac{\varepsilon^{2}}{2}\eta^{2}(x)M(\phi_{1}^{2})\frac{\partial^{2}U}{\partial x^{2}}(t,x)+\varepsilon^{2}o_{R}(1),

where 𝒐𝑹(𝟏)o_{R}(1) tends uniformly towards 𝟎 on [𝟎,𝑻]×[𝑹,𝑹][0,T]\times[-R,R] as 𝜺𝟎\varepsilon\to 0 (the uniformity of the reminder term can be observed with classical computations, let us just note that 𝒐𝑹(𝟏)o_{R}(1) is a generic notation in the sequel). The equation (3.48) and the particular link between both functions 𝑭F and 𝜼\eta imply

𝓜𝜺𝑼(𝒕,𝒙)\displaystyle\mathcal{M}_{\varepsilon}U(t,x) =𝑼(𝒕,𝒙)𝑴(𝒕/(𝜺𝟐𝜼𝟐(𝒙)))𝜺𝟐+𝜺𝟐𝒐𝑹(𝟏),(𝒕,𝒙)[𝟎,𝑻]×[𝑹,𝑹].\displaystyle=U(t,x)M(t/(\varepsilon^{2}\eta^{2}(x)))-\varepsilon^{2}+\varepsilon^{2}o_{R}(1),\quad\forall(t,x)\in[0,T]\times[-R,R]. (3.49)

Step 3. Study of the difference Dε(t,x)D^{\varepsilon}(t,x) introduced in (3.44). Since both 𝒖𝜺u^{\varepsilon} and 𝑼U are continuous functions satisfying an exponential bound (immediate consequence of the regularity and growth property of 𝑭F for 𝑼U and statement of Proposition 3.11 for 𝒖𝜺u^{\varepsilon}), so is 𝑫𝜺D^{\varepsilon}. Hence, there exists 𝑪>𝟎C>0 and 𝜿>𝟎\kappa>0 such that

|𝑫𝜺(𝒕,𝒙)|𝑪𝒆𝜿|𝒙|,(𝒕,𝒙)[𝟎,𝑻]×.\displaystyle|D^{\varepsilon}(t,x)|\leq Ce^{\kappa|x|},\quad\forall(t,x)\in[0,T]\times\mathbb{R}. (3.50)

Moreover combining (3.49) and (3.47) implies

𝑫𝜺(𝒕,𝒙)=(𝑼(𝒕,𝒙)+𝜺𝟐)(𝑴(𝒕/(𝜺𝟐𝜼𝟐(𝒙)))𝟏)+𝓜𝜺𝑫𝜺(𝒕,𝒙)+𝜺𝟐𝒐𝑹(𝟏).\displaystyle D^{\varepsilon}(t,x)=(U(t,x)+\varepsilon^{2})(M(t/(\varepsilon^{2}\eta^{2}(x)))-1)+\mathcal{M}_{\varepsilon}D^{\varepsilon}(t,x)+\varepsilon^{2}o_{R}(1). (3.51)

The support of the distribution associated to 𝑴M is compact. Moreover 𝜼\eta defined in (3.11) is upper-bounded. Consequently there exists 𝝆>𝟎\rho>0 (independent of 𝒙x and 𝜺\varepsilon) such that 𝑴(𝒕/(𝜺𝟐𝜼𝟐(𝒙)))=𝟏M(t/(\varepsilon^{2}\eta^{2}(x)))=1 for all 𝒕𝝆𝜺𝟐t\geq\rho\varepsilon^{2} and 𝒙x\in\mathbb{R}. For small values of 𝒕t, that is 𝒕𝝆𝜺𝟐t\leq\rho\varepsilon^{2}, it suffices to use the regularity of 𝑼U with respect to that variable in order to get a constant 𝑪𝑹>𝟎C_{R}>0 such that |𝑼(𝒕,𝒙)|𝑪𝑹𝜺𝟐|U(t,x)|\leq C_{R}\varepsilon^{2} for all 𝒙[𝑹,𝑹]x\in[-R,R]. To sum up the observations for any value of 𝒕t: there exists 𝑪𝑹>𝟎C_{R}>0 such that

|𝑫𝜺(𝒕,𝒙)|𝓜𝜺|𝑫𝜺|(𝒕,𝒙)+(𝑪𝑹+𝟏)𝜺𝟐𝟏{𝒕𝝆𝜺𝟐}+𝜺𝟐𝒐𝑹(𝟏),(𝒕,𝒙)[𝟎,𝑻]×[𝑹,𝑹].\displaystyle|D^{\varepsilon}(t,x)|\leq\mathcal{M}_{\varepsilon}|D^{\varepsilon}|(t,x)+(C_{R}+1)\,\varepsilon^{2}1_{\{t\leq\rho\varepsilon^{2}\}}+\varepsilon^{2}o_{R}(1),\quad\forall(t,x)\in[0,T]\times[-R,R]. (3.52)

Step 4. Asymptotic behaviour of Dε(t,x)D^{\varepsilon}(t,x). It is possible to link the operator 𝓜𝜺\mathcal{M}_{\varepsilon} to the approximation scheme of the Brownian motion: the Brownian skeleton (𝐁𝐒)𝜼({\rm BS})_{\eta}. We recall that 𝒔𝒏𝜺=𝒌=𝟏𝒏𝑼𝒌𝜺s_{n}^{\varepsilon}=\sum_{k=1}^{n}U_{k}^{\varepsilon} and that (𝒔𝒏𝜺,𝒙𝒏𝜺)𝒏𝟎(s_{n}^{\varepsilon},x_{n}^{\varepsilon})_{n\geq 0} is a skeleton of the Brownian paths: the sequence (Markov chain) has the same distribution than points belonging to a Brownian trajectory. It represents the successive exit times and positions of small ϕ𝜺\phi_{\varepsilon}-domains also called heat-balls, the radius of any heat-ball being upper-bounded by 𝜺𝜼(𝟎)\varepsilon\eta(0). We observe that

𝓜𝜺(𝒉)(𝒕,𝒙)=𝔼[𝒉(𝒕𝑼𝟏𝜺,𝒙𝟏𝜺)𝟏{𝑼𝟏𝜺𝒕}|𝒙𝟎𝜺=𝒙],for any nonnegative function 𝒉.\mathcal{M}_{\varepsilon}(h)(t,x)=\mathbb{E}[h(t-U^{\varepsilon}_{1},x^{\varepsilon}_{1})1_{\{U^{\varepsilon}_{1}\leq t\}}|x_{0}^{\varepsilon}=x],\quad\mbox{for any nonnegative function }h.

Consequently, for any (𝒕,𝒙)[𝟎,𝑻]×[𝑹,𝑹](t,x)\in[0,T]\times[-R,R], (3.52) becomes

|𝑫𝜺(𝒕,𝒙)|𝔼[|𝑫𝜺(𝒕𝑼𝟏𝜺,𝒙𝟏𝜺)|𝟏{𝑼𝟏𝜺𝒕}|𝒙𝟎𝜺=𝒙]+𝑪𝑹𝜺𝟐𝟏{𝒕𝝆𝜺𝟐}+𝜺𝟐𝒐𝑹(𝟏).\displaystyle|D^{\varepsilon}(t,x)|\leq\mathbb{E}[|D^{\varepsilon}(t-U^{\varepsilon}_{1},x^{\varepsilon}_{1})|1_{\{U^{\varepsilon}_{1}\leq t\}}|x_{0}^{\varepsilon}=x]+C_{R}\,\varepsilon^{2}1_{\{t\leq\rho\varepsilon^{2}\}}+\varepsilon^{2}o_{R}(1). (3.53)

Since the sequence (𝒔𝒏𝜺,𝒙𝒏𝜺)𝒏𝟎(s_{n}^{\varepsilon},x_{n}^{\varepsilon})_{n\geq 0} is a Markov chain, the aim is to iterate the upper-bound a large number of times. In order to achieve such a procedure, we need to ensure that 𝒙𝟏𝜺,,𝒙𝒏𝜺x_{1}^{\varepsilon},\ldots,x_{n}^{\varepsilon} belong to the interval [𝑹,𝑹][-R,R]. We introduce

𝝉𝑹,𝜺=𝐢𝐧𝐟{𝒏𝟎:𝒙𝒏𝜺[𝑹,𝑹]}.\tau_{R,\varepsilon}=\inf\{n\geq 0:\ x_{n}^{\varepsilon}\notin[-R,R]\}.

The ϕ𝜺\phi_{\varepsilon}-domains associated to the Brownian approximation are bounded (their radius is less than 𝜺𝜼(𝟎)\varepsilon\eta(0)), we therefore obtain that |𝒙𝝉𝑹,𝜺|𝑹+𝜺𝜼(𝟎)|x_{\tau_{R,\varepsilon}}|\leq R+\varepsilon\eta(0) and (3.50) implies the existence of 𝑪>𝟎C>0 and 𝜿>𝟎\kappa>0 such that |𝑫𝜺(𝒕,𝒙𝝉𝑹,𝜺)|𝑪𝒆𝜿𝑹|D^{\varepsilon}(t,x_{\tau_{R,\varepsilon}})|\leq Ce^{\kappa R}, for any 𝒕𝑻t\leq T and 𝜺𝟏\varepsilon\leq 1. Let us note that for notational convenience we use 𝒙\mathbb{P}_{x} (resp. 𝔼𝒙\mathbb{E}_{x}) for the conditional probability (resp. expectation) with respect to the event 𝒙𝟎𝜺=𝒙x_{0}^{\varepsilon}=x. Hence (3.52) gives

|𝑫𝜺(𝒕,𝒙)|\displaystyle|D^{\varepsilon}(t,x)| 𝔼𝒙[|𝑫𝜺(𝒕𝒔𝟏𝜺,𝒙𝟏𝜺)|𝟏{𝒔𝟏𝜺𝒕;|𝒙𝟏𝜺|𝑹}]+𝔼𝒙[|𝑫𝜺(𝒕𝒔𝟏𝜺,𝒙𝟏𝜺)|𝟏{𝒔𝝉𝑹,𝜺𝜺𝒕;𝝉𝑹,𝜺=𝟏}]\displaystyle\leq\mathbb{E}_{x}[|D^{\varepsilon}(t-s^{\varepsilon}_{1},x^{\varepsilon}_{1})|1_{\{s^{\varepsilon}_{1}\leq t\,;\ |x_{1}^{\varepsilon}|\leq R\}}]+\mathbb{E}_{x}[|D^{\varepsilon}(t-s^{\varepsilon}_{1},x^{\varepsilon}_{1})|1_{\{s^{\varepsilon}_{\tau_{R,\varepsilon}}\leq t\,;\ \tau_{R,\varepsilon}=1\}}]
+𝑪𝑹𝜺𝟐𝟏{𝒕𝝆𝜺𝟐}+𝜺𝟐𝒐𝑹(𝟏)\displaystyle\quad+C_{R}\,\varepsilon^{2}1_{\{t\leq\rho\varepsilon^{2}\}}+\varepsilon^{2}o_{R}(1)
𝔼𝒙[|𝑫𝜺(𝒕𝒔𝟏𝜺,𝒙𝟏𝜺)|𝟏{𝒔𝟏𝜺𝒕;|𝒙𝟏𝜺|𝑹}]+𝑪𝒆𝜿𝑹𝒙(𝒔𝝉𝑹,𝜺𝜺𝒕;𝝉𝑹,𝜺=𝟏)\displaystyle\leq\mathbb{E}_{x}[|D^{\varepsilon}(t-s^{\varepsilon}_{1},x^{\varepsilon}_{1})|1_{\{s^{\varepsilon}_{1}\leq t\,;\ |x_{1}^{\varepsilon}|\leq R\}}]+Ce^{\kappa R}\,\mathbb{P}_{x}(s^{\varepsilon}_{\tau_{R,\varepsilon}}\leq t;\ \tau_{R,\varepsilon}=1)
+𝑪𝑹𝜺𝟐𝟏{𝒕𝝆𝜺𝟐}+𝜺𝟐𝒐𝑹(𝟏).\displaystyle\quad+C_{R}\,\varepsilon^{2}1_{\{t\leq\rho\varepsilon^{2}\}}+\varepsilon^{2}o_{R}(1).

In order to simply the notations when iterating the procedure, we introduce the following events:

𝓙𝑹,𝜺𝒏:={𝒔𝒏𝜺𝒕}{𝝉𝑹,𝜺>𝒏}.\displaystyle\mathcal{J}_{R,\varepsilon}^{n}:=\{s^{\varepsilon}_{n}\leq t\}\cap\{\tau_{R,\varepsilon}>n\}.

By iterating the upper-bound, we obtain

|𝑫𝜺(𝒕,𝒙)|\displaystyle|D^{\varepsilon}(t,x)| 𝔼𝒙[|𝑫𝜺(𝒕𝒔𝟐𝜺,𝒙𝟐𝜺)|𝟏𝓙𝑹,𝜺𝟐]+𝑪𝑹𝜺𝟐𝒙(𝒕𝝆𝜺𝟐𝒔𝟏𝜺𝒕;𝓙𝑹,𝜺𝟐)\displaystyle\leq\mathbb{E}_{x}[|D^{\varepsilon}(t-s^{\varepsilon}_{2},x^{\varepsilon}_{2})|1_{\mathcal{J}_{R,\varepsilon}^{2}}]+C_{R}\,\varepsilon^{2}\,\mathbb{P}_{x}(t-\rho\varepsilon^{2}\leq s_{1}^{\varepsilon}\leq t\,;\,\mathcal{J}_{R,\varepsilon}^{2})
+𝑪𝒆𝜿𝑹𝒙(𝒔𝝉𝑹,𝜺𝜺𝒕;𝝉𝑹,𝜺𝟐)+𝑪𝑹𝜺𝟐𝟏{𝒕𝝆𝜺𝟐}+𝟐𝜺𝟐𝒐𝑹(𝟏)\displaystyle\quad+C\,e^{\kappa R}\,\mathbb{P}_{x}(s^{\varepsilon}_{\tau_{R,\varepsilon}}\leq t;\ \tau_{R,\varepsilon}\leq 2)+C_{R}\,\varepsilon^{2}1_{\{t\leq\rho\varepsilon^{2}\}}+2\,\varepsilon^{2}o_{R}(1)
𝔼𝒙[|𝑫𝜺(𝒕𝒔𝒏𝜺,𝒙𝒏𝜺)|𝟏𝓙𝑹,𝜺𝒏]+𝑪𝑹𝜺𝟐𝒌𝟏𝒙(𝒕𝝆𝜺𝟐𝒔𝒌𝜺𝒕;𝓙𝑹,𝜺𝒌)\displaystyle\leq\mathbb{E}_{x}[|D^{\varepsilon}(t-s^{\varepsilon}_{n},x^{\varepsilon}_{n})|1_{\mathcal{J}_{R,\varepsilon}^{n}}]+C_{R}\,\varepsilon^{2}\sum_{k\geq 1}\mathbb{P}_{x}(t-\rho\varepsilon^{2}\leq s_{k}^{\varepsilon}\leq t\,;\mathcal{J}_{R,\varepsilon}^{k})
+𝑪𝒆𝜿𝑹𝒙(𝒔𝝉𝑹,𝜺𝜺𝒕;𝝉𝑹,𝜺𝒏)+𝑪𝑹𝜺𝟐𝟏{𝒕𝝆𝜺𝟐}+𝒏𝜺𝟐𝒐𝑹(𝟏)\displaystyle\quad+C\,e^{\kappa R}\,\mathbb{P}_{x}(s^{\varepsilon}_{\tau_{R,\varepsilon}}\leq t;\ \tau_{R,\varepsilon}\leq n)+C_{R}\,\varepsilon^{2}1_{\{t\leq\rho\varepsilon^{2}\}}+n\,\varepsilon^{2}o_{R}(1)
𝑪𝑹(𝓐𝟏(𝑹,𝜺,𝒏)+𝓐𝟐(𝑹,𝜺)+𝓐𝟑(𝑹)+𝓐𝟒(𝑹,𝜺,𝒏)),\displaystyle\leq C_{R}\Big{(}\mathcal{A}_{1}(R,\varepsilon,n)+\mathcal{A}_{2}(R,\varepsilon)+\mathcal{A}_{3}(R)+\mathcal{A}_{4}(R,\varepsilon,n)\Big{)}, (3.54)

with

𝓐𝟏(𝑹,𝜺,𝒏):=𝒙(𝓙𝑹,𝜺𝒏),𝓐𝟐(𝑹,𝜺)=𝜺𝟐𝒌𝟏𝒙(𝒕𝝆𝜺𝟐𝒔𝒌𝜺𝒕;𝓙𝑹,𝜺𝒌),𝓐𝟑(𝑹)=𝒆𝜿𝑹(𝒔𝒕:𝒙+𝑩𝒔[𝑹,𝑹]),𝓐𝟒(𝑹,𝜺,𝒏)=𝜺𝟐𝟏{𝒕𝝆𝜺𝟐}+𝒏𝜺𝟐𝒐𝑹(𝟏).\displaystyle\begin{array}[]{ll}\displaystyle\mathcal{A}_{1}(R,\varepsilon,n):=\,\mathbb{P}_{x}(\mathcal{J}_{R,\varepsilon}^{n}),&\displaystyle\mathcal{A}_{2}(R,\varepsilon)=\varepsilon^{2}\sum_{k\geq 1}\mathbb{P}_{x}(t-\rho\varepsilon^{2}\leq s_{k}^{\varepsilon}\leq t\,;\mathcal{J}_{R,\varepsilon}^{k}),\\[5.0pt] \displaystyle\mathcal{A}_{3}(R)=e^{\kappa R}\,\mathbb{P}(\exists s\leq t:\,x+B_{s}\notin[-R,R]),&\mathcal{A}_{4}(R,\varepsilon,n)=\varepsilon^{2}1_{\{t\leq\rho\varepsilon^{2}\}}+n\,\varepsilon^{2}o_{R}(1).\end{array}

We shall now describe precisely the bound of each of these terms. The crucial idea is to first fix 𝑹R sufficiently large and then to choose 𝒏=𝝃𝟏/𝜺𝟐n=\xi\lfloor 1/\varepsilon^{2}\rfloor for 𝝃\xi large enough and depending on 𝑹R. Let 𝜹>𝟎\delta>0. We shall prove that there exists 𝜺𝟎\varepsilon_{0} such that |𝑫𝜺(𝒕,𝒙)|𝜹|D^{\varepsilon}(t,x)|\leq\delta for 𝜺𝜺𝟎\varepsilon\leq\varepsilon_{0}.

  1. 1.

    Due to the reflection principle of the Brownian motion, there exists 𝑹R large enough such that

    𝓐𝟑(𝑹)𝟒𝒆𝜿𝑹(𝑩𝒕𝑹|𝒙|)=𝟒𝒆𝜿𝑹(𝑮𝑹|𝒙|𝒕)𝜹/𝟒,\mathcal{A}_{3}(R)\leq 4e^{\kappa R}\,\mathbb{P}(B_{t}\geq R-|x|)=4e^{\kappa R}\,\mathbb{P}\Big{(}G\geq\frac{R-|x|}{\sqrt{t}}\Big{)}\leq\delta/4, (3.56)

    where 𝑮G is a standard Gaussian variate. From now on, 𝑹R is fixed s.t. (3.56) is satisfied.

  2. 2.

    Let us consider the term 𝓐𝟐\mathcal{A}_{2}. We introduce the particular choice 𝒏=𝝃𝟏/𝜺𝟐n=\xi\lfloor 1/\varepsilon^{2}\rfloor with 𝝃\xi\in\mathbb{N}. By the definition of the Brownian skeleton, 𝒔𝒏𝜺𝒕s_{n}^{\varepsilon}\leq t corresponds to

    𝒌=𝟏𝝃𝟏/𝜺𝟐𝜺𝟐𝜼𝟐(𝒙𝒌𝟏𝜺)𝒆𝟏𝑨𝒌𝒕,\sum_{k=1}^{\xi\lfloor 1/\varepsilon^{2}\rfloor}\varepsilon^{2}\eta^{2}(x_{k-1}^{\varepsilon})e^{1-A_{k}}\leq t,

    where (𝑨𝒌)𝒌𝟏(A_{k})_{k\geq 1} is a sequence of i.i.d random variables. Since 𝜼\eta is an even function and decreases on +\mathbb{R}_{+}, we observe :

    𝓙𝑹,𝜺𝒏{𝜺𝟐𝝃𝒌=𝟏𝝃𝟏/𝜺𝟐𝒆𝟏𝑨𝒌𝒕𝝃𝜼𝟐(𝑹),},𝝃.\mathcal{J}_{R,\varepsilon}^{n}\subset\left\{\frac{\varepsilon^{2}}{\xi}\sum_{k=1}^{\xi\lfloor 1/\varepsilon^{2}\rfloor}e^{1-A_{k}}\leq\frac{t}{\xi\eta^{2}(R)},\right\},\quad\forall\xi\in\mathbb{N}.

    By the law of large numbers, the left hand side of the inequality converges towards 𝔼[𝒆𝟏𝑨𝟏]\mathbb{E}[e^{1-A_{1}}] as 𝜺𝟎\varepsilon\to 0. Hence, as soon as 𝝃>𝒕/(𝜼𝟐(𝑹)𝔼[𝒆𝟏𝑨𝟏])\xi>t/(\eta^{2}(R)\mathbb{E}[e^{1-A_{1}}]), there exists 𝜺𝟏>𝟎\varepsilon_{1}>0 such that 𝓐𝟏(𝑹,𝜺,𝒏)𝜹/𝟒\mathcal{A}_{1}(R,\varepsilon,n)\leq\delta/4 for 𝜺𝜺𝟏\varepsilon\leq\varepsilon_{1} and 𝒏=𝝃𝟏/𝜺𝟐n=\xi\lfloor 1/\varepsilon^{2}\rfloor.

  3. 3.

    Let us now deal with 𝓐𝟒\mathcal{A}_{4}. The parameters 𝑹R and 𝝃\xi have already been fixed and 𝒏=𝝃𝟏/𝜺𝟐n=\xi\lfloor 1/\varepsilon^{2}\rfloor. It is therefore obvious that there exists a constant 𝜺𝟐>𝟎\varepsilon_{2}>0 such that 𝓐𝟒(𝑹,𝜺,𝒏)𝜹/𝟒\mathcal{A}_{4}(R,\varepsilon,n)\leq\delta/4 for 𝜺𝜺𝟐\varepsilon\leq\varepsilon_{2}.

  4. 4.

    Finally we focus our attention on the last term 𝓐𝟐\mathcal{A}_{2} (𝑹R being fixed). We introduce the notation 𝝌(𝑨,𝑩)=𝟏𝑨𝑩\chi(A,B)=1_{A\cap B} and the stopping time

    𝖟=𝐢𝐧𝐟{𝒏𝟎:𝒔𝒏𝜺𝒕𝝆𝜺𝟐}.\mathfrak{z}=\inf\{n\geq 0:\ s_{n}^{\varepsilon}\geq t-\rho\varepsilon^{2}\}.

    Then

    𝜺𝟐𝓐𝟐(𝑹,𝜺)\displaystyle\varepsilon^{-2}\mathcal{A}_{2}(R,\varepsilon) =𝔼[𝒌𝟏𝝌(𝒕𝝆𝜺𝟐𝒔𝒌𝜺𝒕,𝒌<𝝉𝑹,𝜺)]=𝔼[𝒌𝖟𝝌(𝒔𝒌𝜺𝒕,𝒌<𝝉𝑹,𝜺)]\displaystyle=\mathbb{E}\left[\sum_{k\geq 1}\chi(t-\rho\varepsilon^{2}\leq s_{k}^{\varepsilon}\leq t,k<\tau_{R,\varepsilon})\right]=\mathbb{E}\left[\sum_{k\geq\mathfrak{z}}\chi(s_{k}^{\varepsilon}\leq t,k<\tau_{R,\varepsilon})\right]
    𝟏+𝔼[𝒌𝟏𝝌(𝑼𝖟+𝟏𝜺++𝑼𝖟+𝒌𝜺𝒕𝒔𝖟𝜺,𝖟+𝒌<𝝉𝑹,𝜺)].\displaystyle\leq 1+\mathbb{E}\left[\sum_{k\geq 1}\chi(U^{\varepsilon}_{\mathfrak{z}+1}+\ldots+U^{\varepsilon}_{\mathfrak{z}+k}\leq t-s_{\mathfrak{z}}^{\varepsilon},\mathfrak{z}+k<\tau_{R,\varepsilon})\right].

    By definition 𝑼𝒏𝜺=𝜺𝟐𝜼𝟐(𝒙𝒏𝟏𝜺)𝒆𝟏𝑨𝒏𝜺𝟐𝜼𝟐(𝑹)𝒆𝟏𝑨𝒏U^{\varepsilon}_{n}=\varepsilon^{2}\eta^{2}(x_{n-1}^{\varepsilon})e^{1-A_{n}}\geq\varepsilon^{2}\eta^{2}(R)e^{1-A_{n}} for any 𝒏<𝝉𝑹,𝜺n<\tau_{R,\varepsilon}, since 𝜼\eta is decreasing on +\mathbb{R}_{+} and corresponds to an even function. Moreover the definition of 𝖟\mathfrak{z} implies 𝒕𝒔𝖟𝜺𝝆𝜺𝟐t-s_{\mathfrak{z}}^{\varepsilon}\leq\rho\varepsilon^{2}. We deduce that

    𝜺𝟐𝓐𝟐(𝑹,𝜺)\displaystyle\varepsilon^{-2}\mathcal{A}_{2}(R,\varepsilon) 𝟏+𝔼[𝒌𝟏𝝌(𝒆𝟏𝑨𝖟+𝟏++𝒆𝟏𝑨𝖟+𝒌𝝆𝜼𝟐(𝑹),𝖟+𝒌<𝝉𝑹,𝜺)].\displaystyle\leq 1+\mathbb{E}\left[\sum_{k\geq 1}\chi\Big{(}e^{1-A_{\mathfrak{z}+1}}+\ldots+e^{1-A_{\mathfrak{z}+k}}\leq\frac{\rho}{\eta^{2}(R)},\mathfrak{z}+k<\tau_{R,\varepsilon}\Big{)}\right].

    Since (𝒆𝟏𝑨𝒏)𝒏𝟏(e^{1-A_{n}})_{n\geq 1} is a sequence of i.i.d random variables, we can define the associate renewal process (𝑵¯𝒕)𝒕𝟎(\overline{N}_{t})_{t\geq 0} already introduced in the proof of Proposition 2.3. We obtain

    𝜺𝟐𝓐𝟐(𝑹,𝜺)\displaystyle\varepsilon^{-2}\mathcal{A}_{2}(R,\varepsilon) 𝟏+𝔼[𝒌𝟏𝝌(𝒆𝟏𝑨𝟏++𝒆𝟏𝑨𝒌𝝆𝜼𝟐(𝑹))]𝟏+𝔼[𝑵¯𝝆𝜼𝟐(𝑹)]<.\displaystyle\leq 1+\mathbb{E}\left[\sum_{k\geq 1}\chi\Big{(}e^{1-A_{1}}+\ldots+e^{1-A_{k}}\leq\frac{\rho}{\eta^{2}(R)}\Big{)}\right]\leq 1+\mathbb{E}[\overline{N}_{\frac{\rho}{\eta^{2}(R)}}]<\infty.

    In other words, there exists 𝜺𝟑>𝟎\varepsilon_{3}>0 s.t. 𝓐𝟐(𝑹,𝜺)𝜹/𝟒\mathcal{A}_{2}(R,\varepsilon)\leq\delta/4 for any 𝜺𝜺𝟑\varepsilon\leq\varepsilon_{3}.

Let us combine the asymptotic analysis of each term in (3). Then, for any 𝜹>𝟎\delta>0, we define 𝜺𝟎:=𝐦𝐢𝐧(𝜺𝟏,𝜺𝟐,𝜺𝟑)\varepsilon_{0}:=\min(\varepsilon_{1},\varepsilon_{2},\varepsilon_{3}) which insures the announced statement: |𝑫𝜺(𝒕,𝒙)|𝜹|D^{\varepsilon}(t,x)|\leq\delta for any 𝜺𝜺𝟎\varepsilon\leq\varepsilon_{0}. ∎

4 Numerical application

Let us focus our attention on particular examples of LL-class diffusion processes. We recall that these diffusion processes are characterized by their drift term a(t)x+b(t)a(t)x+b(t) and their diffusion coefficient σ¯(t)\overline{\sigma}(t). In many situations, both the particular function f(t,x)f(t,x) and the time scale ρ(t)\rho(t) which permit to write the diffusion process as a function of the time-changed Brownian motion Xt=f(t,x0+Bρ(t))X_{t}=f(t,x_{0}+B_{\rho(t)}) have an explicit formula. We propose two particular cases already introduced in exit problem studies [15].

For each one of these cases we first illustrate some of the results obtained in the theoretical part. Secondly we compare our approach with classical schemes like the Euler scheme. Even if this comparaison is quite difficult as our method looks for a control on the path with an ε\varepsilon approximation while the classical methods do not follow this objective, we construct a rough comparaison that we explain later on.
Example 1 (periodic functions). We set:

a(t)=cos(t)2+sin(t),b(t)=cos(t),andσ¯(t)=2+sin(t).a(t)=\frac{\cos(t)}{2+\sin(t)},\quad b(t)=\cos(t),\quad\mbox{and}\quad\overline{\sigma}(t)=2+\sin(t). (4.1)

Then the three basic components of the ε\varepsilon-strong approximation (see Theorem 3.8) are given by ρ(t)=4t\rho(t)=4t,

f(t,x)=(2+sin(x))(x2+ln(1+sin(t)2))andF(x)=3+|x|2.f(t,x)=(2+\sin(x))\Big{(}\frac{x}{2}+\ln\Big{(}1+\frac{\sin(t)}{2}\Big{)}\Big{)}\quad\mbox{and}\quad F(x)=3+\frac{\sqrt{|x|}}{2}.

We observe that the simulation of a ε\varepsilon-strong approximation of the diffusion paths (Xt,t[0,1])(X_{t},\ t\in[0,1]) requires a random number of ϕε\phi_{\varepsilon}-domains illustrated by the histogram in Figure 1.

Refer to captionRefer to caption

Figure 1: Histogram of the number of ϕε\phi_{\varepsilon}-domains used to cover the time interval [0,1][0,1] for ε=0.1\varepsilon=0.1 (left, sample of size 10 00010\,000) – Average number of ϕε\phi_{\varepsilon}-domains versus the inverse strength 1/ε21/\varepsilon^{2} (right, sample of size for each point: 1 0001\,000). For both pictures: x0=0x_{0}=0.

As said before, it is quite difficult to compare such a method with other numerical approximations of diffusion processes: the other methods don’t lead to build paths which are a.s. ε\varepsilon-close to the diffusion ones. Let us nevertheless sketch a rough comparison: the simulation of 10 00010\,000 paths on [0,1][0,1] with ε=0.1\varepsilon=0.1 requires about 255.7255.7 sec and one can observe that the average time step is about 31053\cdot 10^{-5}. If we consider the classical Euler-scheme with the corresponding constant step size, then a similar sample of paths requires about 41.341.3 sec (on the same computer). One argument which permits to explain the difference in speed is that the ε\varepsilon-strong approximation needs at each step to test if the number of ϕε\phi_{\varepsilon}-domains used so far is sufficient to cover the time interval, such a test is quite time-consuming. Let us also note that the ε\varepsilon-strong approximation permits to be precise not only in the approximation of the marginal distribution but also in the approximation of the whole trajectory. In other words, it is a useful tool for Monte Carlo estimation of an integral, of a supremum, of any functional of the diffusion.

This first example illustrates also the convergence result presented in Theorem 3.13. Since the limiting value is expressed as an average integral of a Brownian motion path, the use of the Monte Carlo procedure permits to get an approximated value: 347.1 on one hand and on the other hand the estimation of the regression line in Figure 1 (right) indicates

(N^1ε)344.3×1ε2+418.5\mathcal{M}(\hat{N}^{\varepsilon}_{1})\approx 344.3\times\frac{1}{\varepsilon^{2}}+418.5

where \mathcal{M} corresponds to the estimated average value for the sample of size 10 00010\,000.

Example 2 (polynomial decrease). We consider on the time interval [0,1][0,1] the ε\varepsilon-strong approximation of the mean reverting diffusion process given by

a(t)=1211+t,b(t)=0,andσ¯(t)=2.a(t)=\frac{1}{2}\frac{1}{1+t},\quad b(t)=0,\quad\mbox{and}\quad\overline{\sigma}(t)=2. (4.2)

Then we obtain the time-scale function ρ(t)=4ln(1+t)\rho(t)=4\ln(1+t) and f(t,x)=x1+t.f(t,x)=x\sqrt{1+t}. We choose therefore F(x)=2+|x|2F(x)=\sqrt{2}+\frac{\sqrt{|x|}}{2}. The number of ϕε\phi_{\varepsilon}-domains is illustrated in Fig 2. The simulation of a sample of trajectories on the time interval [0,1][0,1] of size 10 00010\,000 requires also about 261 sec for the particular choice ε=0.01\varepsilon=0.01 while the classical Euler scheme generated with a comparable step size 3.31053.3\cdot 10^{-5} requires 3030 sec.

Refer to captionRefer to caption

Figure 2: Histogram of the number of ϕε\phi_{\varepsilon}-domains used to cover the time interval [0,1][0,1] for ε=0.01\varepsilon=0.01 (left, sample size: 10 00010\,000) – Average number of ϕε\phi_{\varepsilon}-domains versus the inverse strength 1/ε21/\varepsilon^{2} (right, sample size 1 0001\,000). For both pictures: x0=0x_{0}=0.

Appendix

Let us just present here useful upper-bounds related to the log-gamma distribution.

Lemma 4.1. —

Let α1\alpha\geq 1 and β1\beta\geq 1 and let us assume that WW is a random variable of a log-gamma distribution type. Its probability distribution function is

fW(t):=1Γ(α)βα(lnt)α1t1/β11[0,1](t),t.f_{W}(t):=\frac{1}{\Gamma(\alpha)\beta^{\alpha}}(-\ln t)^{\alpha-1}t^{1/\beta-1}1_{[0,1]}(t),\quad\forall t\in\mathbb{R}.
  • (1)

    Then

    𝔼[eλW]=k0(1)kλkk!(1+kβ)α.\mathbb{E}[e^{-\lambda W}]=\sum_{k\geq 0}\frac{(-1)^{k}\lambda^{k}}{k!(1+k\beta)^{\alpha}}. (4.3)
  • (2)

    In particular, for α=1\alpha=1, we get

    𝔼[eλW]Γ(1/β)βλ1/β.\mathbb{E}[e^{-\lambda W}]\leq\frac{\Gamma(1/\beta)}{\beta\lambda^{1/\beta}}. (4.4)
  • (3)

    In the case: α>1\alpha>1, for any β>β\beta^{\prime}>\beta, we obtain

    𝔼[eλW]ω(α,β,β)1λ1/β,withω(α,β,β)=(α1)α1Γ(1/β)Γ(α)βαeα1(β1β1)α1.\mathbb{E}[e^{-\lambda W}]\leq\omega(\alpha,\beta,\beta^{\prime})\frac{1}{\lambda^{1/\beta^{\prime}}},\quad\mbox{with}\ \omega(\alpha,\beta,\beta^{\prime})=\frac{(\alpha-1)^{\alpha-1}\Gamma(1/\beta^{\prime})}{\Gamma(\alpha)\beta^{\alpha}e^{\alpha-1}(\beta^{-1}-\beta^{\prime-1})^{\alpha-1}}. (4.5)
Proof.

For (1)(1), let us first note that an easy computation leads to the following moments, for any k1k\geq 1:

𝔼[Wk]=(1+kβ)α.\mathbb{E}[W^{k}]=(1+k\beta)^{-\alpha}. (4.6)

After summing over kk (4.6) we deduce the expression of the Laplace transform (4.3).
For (2)(2), let us first consider the particular case: α=1\alpha=1. Using the expression of the PDF and the change of variable u=λxu=\lambda x, we obtain

𝔼[eλW]\displaystyle\mathbb{E}[e^{-\lambda W}] =1β01eλxx1/β1dx=Γ(1/β)βλ1/βIβ(λ),\displaystyle=\frac{1}{\beta}\int_{0}^{1}e^{-\lambda x}x^{1/\beta-1}\,\mathrm{d}x=\frac{\Gamma(1/\beta)}{\beta\lambda^{1/\beta}}\,I_{\beta}(\lambda),

where Iβ(λ):=1Γ(1/β)0λeuu1/β1duI_{\beta}(\lambda):=\frac{1}{\Gamma(1/\beta)}\int_{0}^{\lambda}e^{-u}u^{1/\beta-1}\mathrm{d}u. We observe that λIβ(λ)\lambda\mapsto I_{\beta}(\lambda) is increasing and limλIβ(λ)=1\lim_{\lambda\to\infty}I_{\beta}(\lambda)=1 which leads to (4.4).
For (3)(3), let us now assume that α>1\alpha>1 and consider β>β\beta^{\prime}>\beta. Then

𝔼[eλW]\displaystyle\mathbb{E}[e^{-\lambda W}] =1Γ(α)βα01eλx(x(1/β1/β)/(α1)lnx)α1x1/β1dx\displaystyle=\frac{1}{\Gamma(\alpha)\beta^{\alpha}}\int_{0}^{1}e^{-\lambda x}\Big{(}-x^{(1/\beta-1/\beta^{\prime})/(\alpha-1)}\ln x\Big{)}^{\alpha-1}x^{1/\beta^{\prime}-1}\,\mathrm{d}x
(α1eβα/(α1)(β1β1))α11Γ(α)01eλxx1/β1dx,\displaystyle\leq\left(\frac{\alpha-1}{e\beta^{\alpha/(\alpha-1)}(\beta^{-1}-\beta^{\prime-1})}\right)^{\alpha-1}\frac{1}{\Gamma(\alpha)}\int_{0}^{1}e^{-\lambda x}x^{1/\beta^{\prime}-1}\mathrm{d}x,

since urlnu(re)1-u^{r}\ln u\leq(re)^{-1}. Moreover the change of variable u=λxu=\lambda x leads to

𝔼[eλW]ω(α,β,β)λ1/βIβ(λ).\mathbb{E}[e^{-\lambda W}]\leq\frac{\omega(\alpha,\beta,\beta^{\prime})}{\lambda^{1/\beta^{\prime}}}\,I_{\beta^{\prime}}(\lambda).

The bound Iβ(λ)1I_{\beta^{\prime}}(\lambda)\leq 1 directly leads to (4.5).

References

  • [1] A. Alfonsi. On the discretization schemes for the CIR (and Bessel squared) processes. Monte Carlo Methods Applications, 11 (4):355–384, 2005.
  • [2] A. Beskos, S. Peluchetti and G. Roberts. ε\varepsilon-strong simulation of the Brownian path. Bernoulli, 18 (4):1223–1248, 2012.
  • [3] J. Blanchet, X. Chen and J. Dong. ε\varepsilon-strong simulation for multidimensional stochastic differential equations via rough path analysis. The Annals of Applied Probability, 27 (1):275–336, 2017.
  • [4] Y. Chen, J. Dong and H. Ni ϵ\epsilon-Strong Simulation of Fractional Brownian Motion and Related Stochastic Differential Equations. arXiv 1902.07824, 2019.
  • [5] N. Chen and Z. Huang. Localization and exact simulation of Brownian motion-driven stochastic differential equations. Math.Oper.Res., 38:591–616, 2013.
  • [6] J. Creutzig, T. Müller-Gronbach and K. Ritter Free-knot spline approximation of stochastic processes. Journal of Complexity, 23 (4-6): 867–889, 2017.
  • [7] D.J. Daley and D. Vere-Jones. An Introduction to the Theory of Point Processes. Volume I: Elementary Theory and Methods. Springer, second edition, 2002.
  • [8] L. Devroye Nonuniform Random Variate Generation Springer, New York, 1986.
  • [9] M. Deaconu and S. Herrmann. Simulation of hitting times for Bessel processes with non integer dimension. Bernoulli, 23 (4B):3744–3771, 2017.
  • [10] M. Deaconu, S. Maire and S. Herrmann. The walk on moving spheres: a new tool for simulating Brownian motion’s exit time from a domain. Mathematics and Computers in Simulation, 135:28–38, 2017.
  • [11] M. Deaconu and S. Herrmann. Hitting time for Bessel processes—walk on moving spheres algorithm (WoMS). The Annals of Applied Probability, 23 (6):2259–2289, 2013.
  • [12] L.C. Evans. Partial differential equations. Graduate Studies in Mathematics, Vol. 19, American Mathematical Society, Providence, RI, Second Edition, 2010.
  • [13] M. Hefter, A. Herzwurm and T. Müller-Gronbach. Lower error bounds for strong approximation of scalar SDEs with non-Lipschitzian coefficients. The Annals of Applied Probability, 29 (1):178-216, 2019.
  • [14] H. R. Lerche. Boundary crossing of Brownian motion. Lecture Notes in Statistics, 40, Springer-Verlag, Berlin, 1986.
  • [15] S. Herrmann and N. Massin. Approximation of exit times for one-dimensional linear diffusion processes. Computers & Mathematics with Applications, 80 (6): 1668–1682, 2020.
  • [16] N. Hofmann, T. Müller-Gronbach and K. Ritter. Linear vs standard information for scalar stochastic differential equations. Algorithms and complexity for continuous problems/Algorithms, computational complexity, and models of computation for nonlinear and multivariate problems (Dagstuhl/South Hadley, MA, 2000). J. Complexity, 18 (2):394–414, 2002.
  • [17] M. Hutzenthaler and A. Jentzen. Numerical approximations of stochastic differential equations with non-globally Lipschitz continuous coefficients. Mem. Amer. Math. Soc., 236, no. 1112, 2015.
  • [18] I. Karatzas and Steven E. Shreve. Brownian motion and stochastic calculus, volume 113 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1991.
  • [19] T. Müller-Gronbach. The optimal uniform approximation of systems of stochastic differential equations. Ann. Appl. Probab. 12 (2):664–690, 2002.
  • [20] K. Pötzelberger and L. Wang. Crossing probabilities for diffusions with piecewise continuous boundaries. Methodology and Computing in Applied Probability, 9 (1): 21-40, 2007.
  • [21] D. Revuz and M. Yor. Continuous Martingales and Brownian Motion. Springer Verlag, Berlin, Heidelberg, 1999.