Abstract

We consider the martingale problem related to the solution of an SDE on the line. It is shown that the solution of this martingale problem can be approximated by solutions of the corresponding time-discrete martingale problems under some conditions. This criterion is especially expedient for establishing the convergence of population processes to SDEs. We also show that the criterion yields a weak Euler scheme approximation of SDEs under fairly weak assumptions on the driving force of the approximating processes.

1. Introduction

It is well known that a rescaled version of the classical Galton-Watson process (GWP) with offspring variance 𝜎2 weakly converges to the unique solution of the following one-dimensional stochastic differential equation (SDE):𝑑𝑋𝑑=𝜎|||𝑋𝑑|||π‘‘π‘Šπ‘‘,(1.1)where π‘Š is a one-dimensional Brownian motion (cf. [1]). One might ask whether it is possible to approximate more general SDEs, driven by a Brownian motion, by generalized GWPs. In [2] it will be shown that this is actually possible. In fact, in [2] the solution of the SDE,𝑑𝑋𝑑=𝛿𝑑,𝑋𝑑𝑑𝑑+πœŽπ‘‘,𝑋𝑑|||𝑋𝑑|||π‘‘π‘Šπ‘‘,(1.2)is weakly approximated by two different types of population-size-dependent GWPs (in the sense of [3–6]) with immigration, where 𝛿 and 𝜎 are suitable nonnegative continuous functions on ℝ+×ℝ. Here the methods of [1] do not apply anymore (cf. Section 3). In the present article, we establish a general criterion for the weak approximation of SDEs by discrete-time processes, which is the crux of the analysis of [2].

To be exact, we focus on the following one-dimensional SDE:𝑑𝑋𝑑=𝑏𝑑,𝑋𝑑𝑑𝑑+π‘Žπ‘‘,π‘‹π‘‘ξ‚π‘‘π‘Šπ‘‘,𝑋0=π‘₯0,(1.3)where π‘₯0βˆˆβ„ and π‘Š is a one-dimensional Brownian motion. The coefficients π‘Ž and 𝑏 are continuous functions on ℝ+×ℝ satisfying||||||+||||||ξ‚€||π‘₯||ξ‚π‘Ž(𝑑,π‘₯)𝑏(𝑑,π‘₯)≀𝐾1+βˆ€π‘‘βˆˆβ„+,π‘₯βˆˆβ„,(1.4)for some finite constant 𝐾>0. We assume that SDE (1.3) has a weak solution. It means that there exists a triplet {𝑋;π‘Š;(Ξ©,β„±,(ℱ𝑑),β„™)} where (Ξ©,β„±,(ℱ𝑑),β„™) is a filtered probability space with (ℱ𝑑) satisfying the usual conditions, π‘Š=(π‘Šπ‘‘βˆΆπ‘‘β‰₯0) is an (ℱ𝑑)-Brownian motion, and 𝑋=(π‘‹π‘‘βˆΆπ‘‘β‰₯0) is a real-valued continuous (ℱ𝑑)-adapted process such that β„™-almost surely,𝑋𝑑=π‘₯0+ξ€œπ‘‘0π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚ξ€œπ‘‘π‘Ÿ+𝑑0π‘Žξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚π‘‘π‘Šπ‘Ÿβˆ€π‘‘β‰₯0.(1.5)Here the latter is an ItΜ‚o-integral. Moreover, we require the solution to be weakly unique, which means that any two solutions coincide in law. For instance, the existence of a unique weak solution is implied by Lipschitz continuity of 𝑏 in π‘₯ (uniformly in 𝑑) and|||ξ‚€π‘Ž(𝑑,π‘₯)βˆ’π‘Žπ‘‘,π‘₯ξ…žξ‚|||ξ‚€|||β‰€β„Žπ‘₯βˆ’π‘₯ξ…ž|||ξ‚βˆ€π‘‘βˆˆβ„+,π‘₯,π‘₯ξ…žβˆˆβ„,(1.6)for some strictly increasing β„ŽβˆΆβ„+→ℝ+ with ∫00+β„Žβˆ’2(𝑒)𝑑𝑒=∞. Note that (1.6) and Lipschitz continuity of 𝑏 even imply the existence of a strongly unique strong solution (Yamada-Watanabe criterion [7]). But the notion of strong solutions and strong uniqueness is beyond our interest.

Our starting point is the fact that any weak solution of (1.3) is a solution of the following martingale problem and vice versa (cf. [8, Section 5.4.B], or [9, Theorem 1.27]).

Definition 1.1. A tuple {𝑋;(Ξ©,β„±,(ℱ𝑑),β„™)} is said to be a solution of the (π‘Ž, 𝑏, π‘₯0)-martingale problem if (Ξ©,β„±,(ℱ𝑑),β„™) is a filtered probability space with (ℱ𝑑) satisfying the usual conditions, and 𝑋=(π‘‹π‘‘βˆΆπ‘‘β‰₯0) is a real-valued continuous (ℱ𝑑)-adapted process such that𝑀𝑑=π‘‹π‘‘βˆ’π‘₯0βˆ’ξ€œπ‘‘0π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚π‘‘π‘Ÿ(1.7)provides a (continuous, mean-zero) square-integrable (ℱ𝑑)-martingale with compensatorβŸ¨π‘€βŸ©π‘‘=ξ€œπ‘‘0π‘Ž2ξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚π‘‘π‘Ÿ.(1.8)The solution is said to be unique if any two solutions coincide in law.

In view of the weak equivalence of the SDE to the martingale problem, discrete-time processes solving the discrete analogue (Definition 2.1) of the (π‘Ž,𝑏,π‘₯0)-martingale problem should approximate weakly the unique solution of SDE (1.3). Theorem 2.2 below shows that this is true under an additional assumption on the moments of the increments (condition (2.3)).

Note that the characterization of discrete or continuous population processes as solutions of martingale problems of the form (1.7)-(1.8), (2.1)-(2.2), respectively, is fairly useful and also common (see, e.g., [10–12]). Especially for real-valued discrete-time processes these characterizations are often easy to see, so that, according to the criterion, the only thing to check is condition (2.3). Also note that the conditions of the famous criterion of Stroock and Varadhan for the weak convergence of Markov chains to SDEs [13, Theorem 11.2.3] are different. In particular, in our framework we do not insist on the Markov property of the approximating processes (cf. the discussion at the end of Section 4). Another alternative approach to the discrete-time approximation of SDEs can be found in the seminal paper [14], see also references therein. In [14] general conditions are given, under which the convergence in distribution (π‘Œπ›Ό,𝑍𝛼)β†’(π‘Œ,𝑍) in the cΓ‘dlΓ g space implies convergence in distribution βˆ«π‘Œπ›Όπ‘‘π‘π›Όβ†’βˆ«π‘Œπ‘‘π‘ of the corresponding stochastic integrals in the cΓ‘dlΓ g space.

In Section 3 we will demonstrate that the criterion of Theorem 2.2 yields an easy proof of the convergence result discussed at the beginning of the Introduction. Moreover, in Section 4 we will apply our criterion to obtain a weak Euler scheme approximation of SDEs under fairly weak assumptions on the driving force of the approximating processes.

2. Main Result

We will regard discrete-time processes as continuous-time cΓ‘dlΓ g processes. For this reason we denote by 𝐷(ℝ) the space of cΓ‘dlΓ g functions from ℝ+ to ℝ. We equip 𝐷(ℝ) with the topology generated by the Skohorod convergence on compacts and consider it as a measurable space with respect to its Borel 𝜎-algebra. Moreover, we set π‘‘πœ–π‘›=π‘›πœ– for every π‘›βˆˆβ„•0 and πœ–>0.

For every π›Όβˆˆβ„• we fix some πœ–π›Ό>0 such that πœ–π›Όβ†’0. For the sake of clarity, we also set 𝑑𝛼𝑛=π‘‘πœ–π›Όπ‘›(=π‘›πœ–π›Ό) for all π‘›βˆˆβ„•0. Now suppose that π‘Žπ›Ό and 𝑏𝛼 are measurable functions on ℝ+×ℝ such that β€–π‘Žβˆ’π‘Žπ›Όβ€–βˆž and β€–π‘βˆ’π‘π›Όβ€–βˆž converge to 0 as π›Όβ†’βˆž, where β€–β‹…β€–βˆž is the usual supremum norm. Let (π‘₯𝛼)βŠ‚β„ satisfy π‘₯𝛼→π‘₯0, and suppose that 𝑋𝛼 is a solution of the following (πœ–π›Ό,π‘Žπ›Ό,𝑏𝛼,π‘₯𝛼)-martingale problem for every 𝛼β‰₯1. Here we write 𝑛𝛼(𝑑) for the largest π‘›βˆˆβ„•0 with 𝑑𝛼𝑛≀𝑑.

Definition 2.1. Suppose that 𝑋𝛼=(π‘‹π›Όπ‘‘βˆΆπ‘‘β‰₯0) is a real-valued process on some probability space (Ξ©,β„±,β„™) whose trajectories are constant on the intervals [𝑑𝛼𝑛,𝑑𝛼𝑛+1), π‘›βˆˆβ„•0. Then 𝑋𝛼 is called a solution of the (πœ–π›Ό,π‘Žπ›Ό,𝑏𝛼,π‘₯𝛼)-martingale problem if𝑀𝛼𝑑=π‘‹π›Όπ‘‘βˆ’π‘₯π›Όβˆ’π‘›π›Ό(𝑑)βˆ’1𝑖=0𝑏𝛼𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚πœ–π›Ό(2.1)provides a (zero-mean) square-integrable martingale (with respect to the natural filtration) with compensator𝑀𝛼𝑑=𝑛𝛼(𝑑)βˆ’1𝑖=0π‘Ž2𝛼𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚πœ–π›Ό.(2.2)

The 𝑋𝛼 could be defined on different probability spaces (Ω𝛼,ℱ𝛼,ℙ𝛼). However, we assume without loss of generality that Ω𝛼=𝐷(ℝ), ℱ𝛼=ℬ(𝐷(ℝ)), and 𝑋𝛼 is the coordinate process of ℙ𝛼 (each cΓ‘dlΓ g process induces a corresponding law on 𝐷(ℝ)). We further assume that there are some π‘ž>2 and 𝛿>1 such that𝔼𝛼|||π‘‹π›Όπ‘‘π›Όπ‘›βˆ’π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1|||π‘žξ‚„β‰€πΆπ‘‡ξ‚€1+𝔼𝛼|||π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1|||π‘žπœ–ξ‚„ξ‚π›Ώπ›Ό(2.3)for every 𝛼β‰₯1 and π‘›βˆˆβ„• with 𝑑𝛼𝑛≀𝑇, where 𝐢𝑇>0 is some finite constant that may depend on 𝑇. (By an induction on 𝑛, (2.3) implies immediately that 𝔼𝛼[|𝑋𝛼𝑑𝛼𝑛|π‘ž]<∞ for all 𝛼 and 𝑛. Lemma 5.1 will provide an even stronger statement.) The following theorem shows that 𝑋𝛼 converges in distribution to the unique solution of (1.3).

Theorem 2.2. Suppose SDE (1.3) subject to (1.4) has a unique weak solution, and denote by β„™ the corresponding law on 𝐷(ℝ). Moreover, let ℙ𝛼 be the law (on 𝐷(ℝ)) of 𝑋𝛼 subject to (2.1)–(2.3). Then ℙ𝛼⇒ℙ as π›Όβ†’βˆž.

Here, β‡’ symbolizes weak convergence. The proof of Theorem 2.2 will be carried out in Section 5. The finiteness of the π‘žth moments for some π‘ž>2 is not always necessary, it is true. From time to time the finiteness of the second moments is sufficient. However, for a general statement involving convenient moment conditions as (2.3), a weakening of π‘ž>2 to π‘ž=2 is hardly possible. The assumption π‘ž>2 is common in the theory of functional, time-discrete approximations of SDEs, SDDEs, and SPDEs (see, e.g., [12, 15]).

3. Example 1: Convergence of Rescaled GWP to (1.1)

As a first application of Theorem 2.2, we show that a rescaled GWP weakly converges to Feller's branching diffusion [16], that is, to the solution of SDE (1.1). Lindvall [1] showed this approximation via the convergence of the finite-dimensional distributions, for which the shape of the Laplace transforms of the transition probabilities is essential. Here, we will exploit the martingale property of the Galton-Watson process (with offspring variance 𝜎2). The latter is an β„•0-valued Markov process 𝑍=(π‘π‘›βˆΆπ‘›βˆˆβ„•0) that can be defined recursively as follows. Choose an initial state 𝑍0βˆˆβ„• and set 𝑍𝑛=βˆ‘π‘π‘›βˆ’1𝑖=1π‘π‘›βˆ’1,𝑖 for all 𝑛β‰₯1, where {𝑁𝑛,π‘–βˆΆπ‘›β‰₯0,𝑖β‰₯1} is a family of i.i.d. β„•0-valued random variables with mean 1 and variance 𝜎2. In addition, we require that the fourth moment of 𝑁1,1 is finite. Thereby 𝑍𝑛 has a finite fourth moment for every π‘›βˆˆβ„•0. Actually, in [1] the finiteness of the fourth moments was not required. On the other hand, the methods used there break down when considering a population-size-dependent branching intensity or an additional general immigration into the system. In contrast, the procedure below still works in those cases (cf. [2]).

Setting π‘πœ–π‘‘πœ–π‘›=πœ–π‘π‘› we obtain a rescaled version, π‘πœ–, of 𝑍. Recall π‘‘πœ–π‘›=π‘›πœ–, hence π‘πœ– is a process having πœ–β„•0={0,πœ–,2πœ–,…} as both its index set and its state space. Now pick (πœ–π›Ό)βŠ‚β„+ such that πœ–π›Όβ†’0, and recall our convention 𝑑𝛼𝑛=π‘‘πœ–π›Όπ‘› and that βŒŠπ‘‘βŒ‹πœ– denotes the largest element 𝑠 of πœ–β„•0 with 𝑠≀𝑑. Regard the process π‘πœ–π›Ό as continuous-time process, 𝑋𝛼, by setting 𝑋𝛼𝑑=π‘πœ–π›ΌβŒŠπ‘‘βŒ‹πœ–π›Ό, and suppose that 𝑋𝛼0=⌊π‘₯0βŒ‹πœ–π›Ό. The latter requires that 𝑍0 actually depends on 𝛼. The domain of 𝑋𝛼 is denoted by (Ω𝛼,ℱ𝛼,ℙ𝛼). It is easy to see that 𝑀𝛼 defined in (2.1) provides a (zero-mean) square-integrable martingale. Moreover, the compensator of 𝑀𝛼 is given by βŸ¨π‘€π›ΌβŸ©π‘‘=𝜎2βˆ‘π‘›π›Ό(𝑑)βˆ’1𝑖=0π‘‹π›Όπ‘‘π›Όπ‘–πœ–π›Ό since, in this case,𝔼𝛼𝑀𝛼𝑑𝛼𝑛2βˆ’ξ‚¬π‘€π›Όξ‚­π‘‘π›Όπ‘›ξ‚βˆ’π‘€ξ‚€ξ‚€π›Όπ‘‘π›Όπ‘›βˆ’12βˆ’ξ‚¬π‘€π›Όξ‚­π‘‘π›Όπ‘›βˆ’1ξ‚βˆ£β„±π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚„=0(3.1)can be checked easily with help ofπ”Όπ›Όξ‚ƒπ‘‹π›Όπ‘‘π›Όπ‘›βˆ£π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚„=π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1,π•π‘Žπ‘Ÿπ›Όξ‚ƒπ‘‹π›Όπ‘‘π›Όπ‘›βˆ£π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚„=𝜎2π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1πœ–π›Ό.(3.2)The formulae in (3.2) are immediate consequences of the well-known moment formulae for 𝑍 (see [17, page 6]) and (ℱ𝑋𝛼𝑑𝛼𝑛) denotes the natural filtration induced by 𝑋𝛼. Hence, 𝑋𝛼 solves the (πœ–π›Ό,π‘Ž,𝑏,π‘₯𝛼)-martingale problem of Definition 2.1 with βˆšπ‘Ž(𝑑,π‘₯)=|π‘₯|, 𝑏≑0 and π‘₯𝛼=⌊π‘₯0βŒ‹πœ–π›Ό. It remains to show (2.3). To this end we state the following lemma.

Lemma 3.1. Assume that πœ‰1,πœ‰2,… are independent random variables on some probability space (Ξ©,β„±,β„™) with 𝔼[πœ‰π‘–]=0 and supπ‘–βˆˆβ„•π”Ό[πœ‰4𝑖]<∞. Let 𝜈 be a further random variable on (Ξ©,β„±,β„™) being independent of (πœ‰π‘–), taking values in β„• and satisfying 𝔼[𝜈4]<∞. Then there is some finite constant 𝐢>0, depending only on the second and the fourth moments of the πœ‰π‘–, such that βˆ‘π”Ό[(πœˆπ‘–=1πœ‰π‘–)4]≀𝐢𝔼[𝜈2].

Proof. By the finiteness of the fourth moments the law of total expectation yields𝔼[(πœˆξ“π‘–=1πœ‰π‘–)4]=π‘›π‘›βˆˆβ„•ξ“π‘–1𝑛=1𝑖2𝑛=1𝑖3𝑛=1𝑖4=1π”Όξ‚ƒπœ‰π‘–1πœ‰π‘–2πœ‰π‘–3πœ‰π‘–4ξ‚„β„™[𝜈=𝑛].(3.3)Since the πœ‰π‘– are independent and centered, the summand on the right-hand side might differ from 0 only if either 𝑖1=𝑖2=𝑖3=𝑖4, or 𝑖1=𝑖2 and 𝑖3=𝑖4≠𝑖1, or 𝑖1=𝑖3 and 𝑖2=𝑖4≠𝑖1, or 𝑖1=𝑖4 and 𝑖2=𝑖3≠𝑖1. Hence,𝔼[(πœˆξ“π‘–=1πœ‰π‘–)4]β‰€π‘›βˆˆβ„•ξ€·ξ‚ξ‚†ξ‚€π‘›+3π‘›π‘›βˆ’1)sup𝑖,π‘—βˆˆβ„•π”Όξ‚ƒπœ‰2π‘–πœ‰2𝑗ℙ[𝜈=𝑛]≀4sup𝑖,π‘—βˆˆβ„•π”Όξ‚ƒπœ‰2π‘–πœ‰2π‘—ξ‚„π”Όξ‚ƒπœˆ2ξ‚„.(3.4)This yields the claim of the lemma with 𝐢=4sup𝑖,π‘—βˆˆβ„•π”Ό[πœ‰2π‘–πœ‰2𝑗].

With help of Lemma 3.1 we obtain𝔼𝛼|||π‘‹π›Όπ‘‘π›Όπ‘›βˆ’π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1|||4ξ‚„=𝔼𝛼[|πœ–π›Όβˆ’1π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1𝑖=1ξ‚€πœ–π›Όπ‘π‘›βˆ’1,π‘–βˆ’πœ–π›Όξ‚|4]=𝔼𝛼[|πœ–π›Όβˆ’1π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1𝑖=1ξ‚€π‘π‘›βˆ’1,𝑖|βˆ’14]πœ–4π›Όβ‰€πΆπ”Όπ›Όπœ–ξ‚ƒξ‚€π›Όβˆ’1π‘‹π›Όπ‘‘π›Όπ‘›βˆ’12ξ‚„πœ–4𝛼≀𝐢1+π”Όπ›Όπ‘‹ξ‚ƒξ‚€π›Όπ‘‘π›Όπ‘›βˆ’14πœ–ξ‚„ξ‚2𝛼(3.5)for some suitable constant 𝐢>0. This shows that (2.3) holds too. Hence, the assumptions of Theorem 2.2 are fulfilled, and the theorem implies that 𝑋𝛼 converges in distribution to the unique solution of (1.1).

4. Example 2: Weak Euler Scheme Approximation of (1.3)

As a second application of Theorem 2.2, we establish a weak Euler scheme approximation of SDE (1.3). Our assumptions are partially weaker than the assumptions of classical results on weak functional Euler scheme approximations. A standard reference for Euler schemes is the monograph [18]; see also references therein. As before we suppose that π‘Ž and 𝑏 are continuous functions on ℝ+×ℝ satisfying (1.4), and that SDE (1.3) possesses a unique weak solution. Now let πœ–>0, recall the notation introduced in Section 2, and consider the following stochastic difference equation (weak Euler scheme):π‘‹πœ–π‘‘πœ–π‘›βˆ’π‘‹πœ–π‘‘πœ–π‘›βˆ’1𝑑=π‘πœ–π‘›βˆ’1,π‘‹πœ–π‘‘πœ–π‘›βˆ’1ξ‚ξ‚€π‘‘πœ–+π‘Žπœ–π‘›βˆ’1,π‘‹πœ–π‘‘πœ–π‘›βˆ’1ξ‚π‘‰πœ–π‘‘πœ–π‘›,π‘‹πœ–π‘‘πœ–0=π‘₯πœ–.(4.1)Here, (π‘₯πœ–) is a sequence in ℝ satisfying π‘₯πœ–β†’π‘₯0 as πœ–β†’0, and π‘‰πœ–={π‘‰πœ–π‘‘πœ–π‘›βˆΆπ‘›βˆˆβ„•} is a family of independent centered random variables with variance πœ– and π”Όπœ–[|π‘‰πœ–π‘‘πœ–π‘›|π‘ž]β‰€πΆπœ–π‘ž/2 for all π‘›βˆˆβ„•, πœ–βˆˆ(0,1], some π‘ž>2, and some finite constant 𝐢>0, where (Ξ©πœ–,β„±πœ–,β„™πœ–) denotes the domain of π‘‰πœ–. For instance, one may set π‘‰πœ–π‘‘πœ–π‘›=βˆšπœ–πœ‰π‘› where {πœ‰π‘›βˆΆπ‘›βˆˆβ„•} is a family of independent centered random variables with variance 1 and the π‘žth moment being bounded uniformly in 𝑛. Note that we do not require that the random variables {π‘‰πœ–π‘‘πœ–π‘›βˆΆπ‘›βˆˆβ„•} are identically distributed. Below we will see that the independence is necessary neither.

By virtue of (1.4), π‘‹πœ–π‘‘πœ–π‘› has a finite π‘žth moment if π‘‹πœ–π‘‘πœ–π‘›βˆ’1 has. It follows by induction that the solution π‘‹πœ–=(π‘‹πœ–π‘‘πœ–π‘›βˆΆπ‘›βˆˆβ„•0) of (4.1) is π‘ž-integrable, and hence square integrable. Equation (4.1) is obviously equivalent to the stochastic sum equationπ‘‹πœ–π‘‘πœ–π‘›=π‘₯πœ–+π‘›βˆ’1𝑖=0π‘ξ‚€π‘‘πœ–π‘–,π‘‹πœ–π‘‘πœ–π‘–ξ‚πœ–+π‘›βˆ’1𝑖=0π‘Žξ‚€π‘‘πœ–π‘–,π‘‹πœ–π‘‘πœ–π‘–ξ‚π‘‰πœ–π‘‘πœ–π‘–+1.(4.2)Suppose that (πœ–π›Ό) is an arbitrary sequence with πœ–π›Όβˆˆ(0,1] and πœ–π›Όβ†’0, set π‘₯𝛼=π‘₯πœ–π›Ό and recall our convention 𝔼𝛼=π”Όπœ–π›Ό, π‘‹π›Όβ‰‘π‘‹πœ–π›Ό, 𝑑𝛼𝑛=π‘‘πœ–π›Όπ‘›. Then it is easy to see that 𝑀𝛼 defined in (2.1) provides a (mean-zero) square-integrable (ℱ𝑋𝛼𝑑)-martingale. Moreover, 𝑀𝛼𝑑𝛼𝑛 coincides with the second sum on the right-hand side of (4.2). Therefore, we also obtain𝑀𝛼𝑑𝛼𝑛=𝑛𝑖=1π”Όπ›Όπ‘Žξ‚€π‘‘ξ‚ƒξ‚€π›Όπ‘–βˆ’1,π‘‹π›Όπ‘‘π›Όπ‘–βˆ’1ξ‚π‘‰πœ–π›Όπ‘‘π›Όπ‘–ξ‚2βˆ£β„±π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚„=π‘›βˆ’1𝑖=0π‘Ž2𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚π”Όπ›Όπ‘‰ξ‚ƒξ‚€πœ–π›Όπ‘‘π›Όπ‘–+12ξ‚„=π‘›βˆ’1𝑖=0π‘Ž2𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚πœ–π›Ό(4.3)which shows that 𝑋𝛼 solves the (πœ–π›Ό,π‘Ž,𝑏,π‘₯𝛼)-martingale problem of Definition 2.1. For an application of Theorem 2.2 it thus remains to show (2.3). But (2.3) follows from𝔼𝛼|||π‘‹π›Όπ‘‘π›Όπ‘›βˆ’π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1|||π‘žξ‚„β‰€2π‘žβˆ’1𝔼𝛼|||π‘ξ‚€π‘‘π›Όπ‘›βˆ’1,π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚πœ–π›Ό|||π‘žξ‚„+𝔼𝛼|||π‘Žξ‚€π‘‘π›Όπ‘›βˆ’1,π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1|||π‘žξ€»π”Όπ›Όξ‚ƒ|||π‘‰πœ–π›Όπ‘‘π›Όπ‘›|||π‘žξ€»ξ‚‡β‰€2π‘žβˆ’1𝐾2π‘žβˆ’1ξ‚€1+𝔼𝛼|||π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1|||π‘žπœ–ξ‚„ξ‚π‘žπ›Ό+𝐾2π‘žβˆ’1ξ‚€1+𝔼𝛼|||π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1|||π‘žξ€»ξ‚πΆπœ–π›Όπ‘ž/2(4.4)for which we used (4.1), the independence of π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1 of π‘‰πœ–π›Ό, (1.4), and 𝔼𝛼[|π‘‰πœ–π›Όπ‘‘π›Όπ‘›|π‘ž]β‰€πΆπœ–π›Όπ‘ž/2. Hence, Theorem 2.2 ensures that 𝑋𝛼 converges in distribution to the unique solution of SDE (1.3).

As mentioned above, the independence of the random variables {π‘‰πœ–π‘‘πœ–π‘›βˆΆπ‘›βˆˆβ„•} is not necessary. The independence was used for (4.3), (4.4), and the martingale property of 𝑀𝛼. But these relations may be valid even if the π‘‰πœ–π‘‘πœ–π‘› are not independent. For instance, let {πœ‰π‘›(𝑖)βˆΆπ‘›,π‘–βˆˆβ„•} be an array of independent centered random variables with variance 1 and π‘žth moments being bounded above by some 𝐢>0 uniformly in 𝑛,𝑖, for some π‘ž>2. Then the martingale property of 𝑀𝛼 and the main statements of (4.3) and (4.4) remain true for π‘‰πœ–π‘‘πœ–1=βˆšπœ–πœ‰1(1) and π‘‰πœ–π‘‘πœ–π‘›=βˆšπœ–πœ‰π‘›(𝑓𝑛(π‘‰πœ–π‘‘πœ–1,…,π‘‰πœ–π‘‘πœ–π‘›βˆ’1)), 𝑛β‰₯2, where 𝑓𝑛 is any measurable mapping from β„π‘›βˆ’1 to β„•. This follows from the following relations which can be shown easily with help of the functional representation theorem for conditional expectations respectively by conditioningπ”Όπ›Όξ‚ƒπ‘‰πœ–π›Όπ‘‘π›Όπ‘›βˆ£β„±π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚„=0,π”Όπ›Όπ‘‰ξ‚ƒξ‚€πœ–π›Όπ‘‘π›Όπ‘–+12βˆ£β„±π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚„=πœ–π›Όπ”Ό,1β‰€π‘–β‰€π‘›βˆ’1,𝛼|||π‘Žξ‚€π‘‘π›Όπ‘›βˆ’1,π‘‹π›Όπ‘‘π›Όπ‘›βˆ’1ξ‚π‘‰πœ–π›Όπ‘‘π›Όπ‘›|||π‘žξ‚„β‰€πΆπœ–π›Όπ‘ž/2.(4.5)If the πœ‰π‘›(𝑖) are not identically distributed, then the π‘‰πœ–π‘‘πœ–π‘› are typically not independent. In particular, the approximating process π‘‹πœ– may be non-Markovian.

5. Proof of Theorem 2.2

Theorem 2.2 is an immediate consequence of Propositions 5.2, 5.5, and the weak equivalence of the martingale problem to the SDE. For the proofs of the two propositions we note that there exist πΎξ…ž>0 and 𝛼0β‰₯1 such that for all 𝛼β‰₯𝛼0, 𝑑β‰₯0, and π‘₯βˆˆβ„,|||π‘Žπ›Ό|||+|||𝑏(𝑑,π‘₯)𝛼|||(𝑑,π‘₯)β‰€πΎξ…žξ‚€||π‘₯||1+.(5.1)This is true since we assumed (1.4) and uniform convergence of π‘Žπ›Ό and 𝑏𝛼 to the coefficients π‘Ž and 𝑏, respectively. Throughout this section we will frequently use the well-known inequality |βˆ‘π‘šπ‘–=1𝑦𝑖|π‘β‰€π‘šπ‘βˆ’1βˆ‘π‘šπ‘–=1|𝑦𝑖|𝑝 for all π‘šβˆˆβ„•, 𝑝β‰₯1 and 𝑦1,…,π‘¦π‘šβˆˆβ„. As a first consequence of (5.1) we obtain Lemma 5.1. For every π‘₯βˆˆβ„+ we write ⌊π‘₯βŒ‹πœ– for the largest element of πœ–β„•0={0,πœ–,2πœ–,…} which is smaller than or equal to π‘₯. Moreover, we assume without loss of generality that πœ–π›Όβ‰€1.

Lemma 5.1. For π‘ž>2 and 𝛿>1 satisfying (2.3) and every 𝑇>0, sup𝛼β‰₯𝛼0𝔼𝛼sup𝑑≀𝑇|||𝑋𝛼𝑑|||π‘žξ‚„<∞.(5.2)

Proof. First of all, note that for the proof it actually suffices to require π‘žβ‰₯2 and 𝛿β‰₯1. Set 𝑆=sup𝛼β‰₯𝛼0|π‘₯𝛼|π‘ž and 𝑆𝛼𝑑=𝔼𝛼[max1≀𝑖≀𝑛𝛼(𝑑)|π‘€π›Όπ‘‘π›Όπ‘–βˆ’π‘€π›Όπ‘‘π›Όπ‘–βˆ’1|π‘ž]. Using Proposition A.1 in the appendix and (5.1) we obtain, for all 𝑑>0 and 𝛼β‰₯𝛼0,𝔼𝛼sup𝑖≀𝑛𝛼(𝑑)|||𝑋𝛼𝑑𝑖|||π‘žξ‚„β‰€3π‘žβˆ’1{𝔼𝛼sup𝑖≀𝑛𝛼(𝑑)|||𝑀𝛼𝑑𝑖|||π‘žξ‚„+𝑆+𝔼𝛼[(𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0|||𝑏𝛼𝑑𝛼𝑖,𝑋𝛼𝑑𝛼𝑖|||πœ–π›Ό)π‘ž]}≀3π‘žβˆ’1πΆπ‘ž{𝔼𝛼[|𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0π‘Ž2𝛼𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚πœ–π›Ό|π‘ž/2]+𝑆𝛼𝑑+𝑆+𝔼𝛼[[𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0|𝑏𝛼𝑑𝛼𝑖,𝑋𝛼𝑑𝛼𝑖|πœ–π›Ό]π‘ž]}β‰€π‘˜π‘ž{𝔼𝛼[[𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0ξ‚€πΎξ…žξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||2πœ–π›Ό]π‘ž/2]+𝑆𝛼𝑑+𝑆+𝔼𝛼[[𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0πΎξ…žξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||ξ‚πœ–π›Ό]π‘ž]},(5.3)where πΆπ‘ž is independent of 𝑑 and 𝛼, and π‘˜π‘ž=3π‘žβˆ’1πΆπ‘ž. By HΓΆlder's inequality we get𝔼𝛼[[𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0ξ‚€πΎξ…žξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||2πœ–π›Ό]π‘ž/2]≀𝔼𝛼[(𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0ξ‚€2πΎξ…ž2ξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||2ξ‚ξ‚π‘ž/2)(𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=0πœ–π›Ό(π‘ž/2)/(π‘ž/2βˆ’1))π‘ž/2βˆ’1]≀𝔼𝛼[(𝑛𝛼(𝑑)βˆ’1βˆ‘π‘–=02π‘ž/2βˆ’1ξ‚€2πΎξ…ž2ξ‚π‘ž/2ξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||π‘žξ‚)𝑛𝛼(𝑑)π‘ž/2βˆ’1πœ–π›Όπ‘ž/2]β‰€π‘π‘žπ‘‘π‘ž/2+π‘π‘žπ‘‘π‘›π‘ž/2βˆ’1𝛼(𝑑)βˆ’1βˆ‘π‘–=0𝔼𝛼sup𝑗≀𝑖|||𝑋𝛼𝑑𝛼𝑗|||π‘žξ‚„πœ–π›Ό,(5.4)where π‘π‘ž=2π‘ž/2βˆ’1(2πΎξ…ž2)π‘ž/2. Analogously, with π‘π‘ž=2π‘žβˆ’1πΎξ…žπ‘ž,𝔼𝛼[[𝑛𝛼(𝑑)βˆ’1𝑖=0πΎξ…žξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||ξ‚πœ–π›Ό]π‘ž]β‰€π‘π‘žπ‘‘π‘ž+π‘π‘žπ‘‘π‘›π‘žβˆ’1𝛼(𝑑)βˆ’1𝑖=0𝔼𝛼sup𝑗≀𝑖|||𝑋𝛼𝑑𝛼𝑗|||π‘žξ‚„πœ–π›Ό.(5.5)Moreover, by (2.3) and (5.1) we obtain, for all 𝑑≀𝑇 and 𝛼β‰₯𝛼0,𝑆𝛼𝑑≀𝑛𝛼(𝑑)𝑖=1𝔼𝛼|||π‘€π›Όπ‘‘π›Όπ‘–βˆ’π‘€π›Όπ‘‘π›Όπ‘–βˆ’1|||π‘žξ‚„β‰€2π‘›π‘žβˆ’1𝛼(𝑑)𝑖=1𝔼𝛼|||π‘‹π›Όπ‘‘π›Όπ‘–βˆ’π‘‹π›Όπ‘‘π›Όπ‘–βˆ’1|||π‘ž+|||π‘π›Όξ‚€π‘‘π›Όπ‘–βˆ’1,π‘‹π›Όπ‘‘π›Όπ‘–βˆ’1|||π‘žπœ–π‘žπ›Όξ‚„β‰€2π‘›π‘žβˆ’1𝛼(𝑑)βˆ’1𝑖=0𝐢𝑇1+𝔼𝛼|||𝑋𝛼𝑑𝛼𝑖|||π‘žπœ–ξ‚„ξ‚π›Ώπ›Ό+π”Όπ›Όξ‚ƒπΎξ…žξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||π‘žπœ–ξ‚ξ‚„π‘žπ›Όξ‚‡β‰€π‘π‘ž,𝑇𝑑+π‘π‘›π‘ž,𝑇𝛼(𝑑)βˆ’1𝑖=0𝔼𝛼sup𝑗≀𝑖|||𝑋𝛼𝑑𝛼𝑗|||π‘žξ‚„πœ–π›Ό,(5.6)where π‘π‘ž,𝑇=2π‘žβˆ’1(𝐢𝑇+πΎξ…ž). By all account we have, for all 𝑑≀𝑇 and 𝛼β‰₯𝛼0,𝔼𝛼sup𝑖≀𝑛𝛼(𝑑)|||𝑋𝛼𝑑𝑖|||π‘žξ‚„β‰€π‘˜π‘žξ‚€π‘{𝑆+π‘ž+π‘π‘ž+π‘π‘ž,π‘‡π‘‘ξ‚ξ‚€π‘žβˆ’1ξ‚βˆ¨1(1+πœ–π›Όπ‘›π›Ό(𝑑)βˆ’1𝑖=0𝔼𝛼sup𝑗≀𝑖||𝑋𝛼𝑑𝛼𝑗||π‘žξ‚„β‰€ξ‚€π‘˜)}π‘žπ‘†+πΆπ‘ž,𝑇+πΆπ‘ž,π‘‡πœ–π›Όπ‘›π›Ό(𝑑)βˆ’1𝑖=0𝔼𝛼sup𝑗≀𝑖||𝑋𝛼𝑑𝛼𝑗||π‘žξ‚„,(5.7)where πΆπ‘ž,𝑇=π‘˜π‘ž(π‘π‘ž+π‘π‘ž+π‘π‘ž,𝑇)(π‘‡π‘žβˆ’1∨1). An application of Lemma A.2 yields𝔼𝛼sup𝑠≀𝑑||𝑋𝛼𝑠||π‘žξ‚„=𝔼𝛼sup𝑖≀𝑛𝛼(𝑑)||𝑋𝛼𝑑𝑖||π‘žξ‚„β‰€ξ‚€π‘˜π‘žπ‘†+πΆπ‘ž,𝑇1+πΆπ‘ž,π‘‡πœ–π›Όξ‚π‘›π›Ό(𝑑)+ξ‚€πΆπ‘ž,π‘‡πœ–π›Όξ‚π‘›π›Ό(𝑑)𝑆,(5.8)where we emphasize that the constants π‘˜π‘ž, 𝑆, and πΆπ‘ž,𝑇 are independent of 𝑑≀𝑇 and 𝛼β‰₯𝛼0. This proves Lemma 5.1 since limsupπ›Όβ†’βˆž(1+πΆπ‘ž,π‘‡πœ–π›Ό)𝑛𝛼(𝑑) is bounded by exp(π‘‘πΆπ‘ž,𝑇) (note that 𝑛𝛼(𝑑)=βŒŠπ‘‘/πœ–π›ΌβŒ‹1≀𝑑/πœ–π›Ό).

Proposition 5.2. If (ℙ𝛼) is tight, then the coordinate process of any weak limit point, that has no mass outside of 𝐢(ℝ), is a solution of the (π‘Ž,𝑏,π‘₯0)-martingale problem of Definition 1.1.

Proof. We consider a weakly convergent subsequence whose limit, β„™, has no mass outside of 𝐢(ℝ). By an abuse of notation, we denote this subsequence by (ℙ𝛼) either. We further write 𝑋 for the coordinate process of β„™. Since 𝑋 is β„™-almost surely continuous, we know [19, Theorem 3.7.8] thatβ„™π›Όβˆ˜πœ‹π‘‘βˆ’11,…,π‘‘π‘˜β‡’β„™βˆ˜πœ‹π‘‘βˆ’11,…,π‘‘π‘˜(5.9)for all 𝑑1,…,π‘‘π‘˜βˆˆβ„+, where πœ‹π‘‘1,…,π‘‘π‘˜βˆΆπ·(ℝ)β†’β„π‘˜ is the usual coordinate projection. In the remainder of the proof we will show in three steps that 𝑀 defined in (1.7) is square-integrable, provides an (ℱ𝑋𝑑)-martingale and has βŸ¨π‘€βŸ© defined in (1.8) as compensator. Here, (ℱ𝑋𝑑) denotes the natural augmentation of the filtration (ℱ𝑋𝑑) induced by 𝑋.
Step 1. With help of Fatou's lemma as well as (5.9) and (5.2) we obtain, for every 𝑇>0,sup𝑑≀𝑇𝔼|||𝑋𝑑|||π‘žξ‚„β‰€sup𝑑≀𝑇liminfπ‘β†’βˆžlimπ›Όβ†’βˆžπ”Όπ›Όξ‚ƒ|||𝑋𝛼𝑑|||π‘žξ—π‘ξ‚„β‰€sup𝑑≀𝑇sup𝛼β‰₯𝛼0𝔼𝛼|||𝑋𝛼𝑑|||π‘žξ‚„<∞.(5.10)Taking (1.4) into account we conclude that 𝑀 defined in (1.7) is square-integrable.Step 2. We next show that 𝑀 is an (ℱ𝑋𝑑)-martingale. It suffices to show that 𝑀 is an (ℱ𝑋𝑑)-martingale; see [20, page 75]. The latter is true if and only if𝔼𝑋𝑑+π‘ βˆ’π‘‹π‘‘βˆ’ξ€œπ‘‘π‘‘+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚ξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„=0(5.11)holds for all 0≀𝑑1<⋯≀𝑑𝑙≀𝑑, 𝑠β‰₯0, 𝑙β‰₯1 and bounded β„Ž1,…,β„Žπ‘™βˆˆπΆ(ℝ) (do not confuse 𝑑𝑖 and 𝑑𝛼𝑖). Since 𝑋𝛼 solves the (πœ–π›Ό,π‘Žπ›Ό,𝑏𝛼,π‘₯𝛼)-martingale problem, we have𝔼𝛼[(𝑋𝛼𝑑+π‘ βˆ’π‘‹π›Όπ‘‘βˆ’π‘›π›Ό(𝑑+𝑠)βˆ’1𝑖=𝑛𝛼(𝑑)𝑏𝛼𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚πœ–π›Ό)𝑙𝑖=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚]=0.(5.12)We are going to verify (5.11) by showing that the left-hand side of (5.12) converges to the left-hand side of (5.11) as π›Όβ†’βˆž. We begin with provinglimπ›Όβ†’βˆžπ”Όπ›Όξ‚ƒπ‘‹π›Όπ‘’π‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ƒπ‘‹ξ‚ξ‚„=𝔼𝑒𝑙𝑖=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„(5.13)for every 𝑒β‰₯0, which together with (5.19) below implies the required convergence. To this end we set π‘₯(𝑁)⋀𝑁=(βˆ’π‘βˆ¨π‘₯) for all π‘₯βˆˆβ„ and 𝑁>0. The right-hand side of|||𝔼𝛼𝑋𝑒𝑙𝛼,(𝑁)𝑖=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚„βˆ’π”Όπ›Όξ‚ƒπ‘‹π›Όπ‘’π‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–|||≀𝔼𝛼|||𝑋𝑒𝛼,(𝑁)βˆ’π‘‹π›Όπ‘’|||𝑙𝑖=1β€–β€–β„Žπ‘–β€–β€–βˆžξ‚„(5.14)can be estimated, for every 𝑇β‰₯𝑒, bysupπ‘Ÿβ‰€π‘‡sup𝛼′β‰₯𝛼0𝔼𝛼′|||π‘‹π›Όβ€²π‘Ÿ|||1||π‘‹π›Όβ€²π‘Ÿ||>𝑁𝑙𝑖=1β€–β€–β„Žπ‘–β€–β€–βˆž(5.15)which tends to 0 as π‘β†’βˆž since {π‘‹π›Όβ€²π‘ŸβˆΆπ‘Ÿβ‰€π‘‡,π›Όξ…žβ‰₯1} is uniformly integrable by (5.2). Therefore, we havelimπ‘β†’βˆžπ”Όπ›Όξ‚ƒπ‘‹π‘’π‘™π›Ό,(𝑁)𝑖=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚„=𝔼𝛼𝑋𝛼𝑒𝑙𝑖=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚„uniformlyin𝛼β‰₯𝛼0(5.16)(and uniformly in 𝑒≀𝑇, for every 𝑇>0). By (5.9) we further obtain for every 𝑁>0,limπ›Όβ†’βˆžπ”Όπ›Όξ‚ƒπ‘‹π‘’π‘™π›Ό,(𝑁)𝑖=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ƒπ‘‹ξ‚ξ‚„=𝔼𝑒𝑙(𝑁)𝑖=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„(5.17)since the mapping (π‘₯1,…,π‘₯𝑙+1)↦π‘₯(𝑁)𝑙+1βˆπ‘™π‘–=1β„Žπ‘–(π‘₯𝑖) from ℝ𝑙+1 to ℝ is bounded and continuous. This is the reason why we introduced the truncation π‘₯(𝑁). By virtue of (5.10), an application of the dominated convergence theorem giveslimπ‘β†’βˆžπ”Όξ‚ƒπ‘‹π‘’π‘™(𝑁)𝑖=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ƒπ‘‹ξ‚ξ‚„=𝔼𝑒𝑙𝑖=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„(5.18)which along with (5.16) and (5.17) implies (5.13). It remains to showlimπ›Όβ†’βˆžπ”Όπ›Ό[𝑛𝛼(𝑑+𝑠)βˆ’1𝑖=𝑛𝛼(𝑑)𝑏𝛼𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚πœ–π›Όπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚ƒξ€œ]=𝔼𝑑𝑑+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„.(5.19)Taking (5.1) and (𝑛𝛼(𝑑+𝑠)βˆ’π‘›π›Ό(𝑑))πœ–π›Όβ‰€π‘ +πœ–π›Ό into account we obtain, analogously to (5.16) and (5.18),limπ‘β†’βˆžπ”Όπ›Ό[𝑛𝛼(𝑑+𝑠)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑)𝑏𝛼𝑑𝛼𝑖,𝑋𝑑𝛼,(𝑁)π›Όπ‘–ξ‚πœ–π›Όπ‘™βˆπ‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚]=𝔼𝛼[𝑛𝛼(𝑑+𝑠)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑)𝑏𝛼𝑑𝛼𝑖,π‘‹π›Όπ‘‘π›Όπ‘–ξ‚πœ–π›Όπ‘™βˆπ‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚]uniformlyin𝛼β‰₯𝛼0,(5.20)respectively,limπ‘β†’βˆžπ”Όξ‚ƒξ€œπ‘‘π‘‘+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿ(𝑁)ξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ƒξ€œξ‚ξ‚„=𝔼𝑑𝑑+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„.(5.21)By the uniform convergence of 𝑏𝛼 to 𝑏 and (𝑛𝛼(𝑑+𝑠)βˆ’π‘›π›Ό(𝑑))πœ–π›Όβ‰€π‘ +πœ–π›Ό, we also have𝔼𝛼[𝑛𝛼(𝑑+𝑠)βˆ’1𝑖=𝑛𝛼(𝑑)𝑏𝛼𝑑𝛼𝑖,𝑋𝑑𝛼,(𝑁)π›Όπ‘–ξ‚πœ–π›Όπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚„=𝔼𝛼[𝑛𝛼(𝑑+𝑠)βˆ’1𝑖=𝑛𝛼(𝑑)𝑏𝑑𝛼𝑖,𝑋𝑑𝛼,(𝑁)π›Όπ‘–ξ‚πœ–π›Όπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚„+π‘œπ›Ό(1).(5.22)Moreover, we have𝔼𝛼[𝑛𝛼(𝑑+𝑠)βˆ’1𝑖=𝑛𝛼(𝑑)𝑏𝑑𝛼𝑖,𝑋𝑑𝛼,(𝑁)π›Όπ‘–ξ‚πœ–π›Όπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚„=π”Όπ›Όξ‚ƒξ€œπ‘‘π‘‘+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿπ›Ό,(𝑁)ξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚„+π‘œπ›Ό(1)(5.23)which is a consequence of the dominated convergence theorem and|𝑛𝛼(𝑑+𝑠)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑)𝑏𝑑𝛼𝑖,𝑋𝑑𝛼,(𝑁)π›Όπ‘–ξ‚πœ–π›Όβˆ’βˆ«π‘‘π‘‘+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿπ›Ό,(𝑁)ξ‚β‰€βˆ«βŒŠπ‘‘π‘Ÿ|𝑑+π‘ βŒ‹πœ–π›Όβˆ’πœ–π›ΌβŒŠπ‘‘βŒ‹πœ–π›Ό|||π‘ξ‚€βŒŠπ‘ŸβŒ‹πœ–π›Ό,π‘‹π‘Ÿπ›Ό,(𝑁)ξ‚ξ‚€βˆ’π‘π‘Ÿ,π‘‹π‘Ÿπ›Ό,(𝑁)|||π‘‘π‘Ÿ+π‘œπ›Ό(1)(5.24)together with the fact that 𝑏 is bounded and uniformly continuous on [0,𝑑+𝑠]Γ—[βˆ’π‘,𝑁]. Finally, we get by (5.9) and the dominated convergence theorem and (5.2) limπ›Όβ†’βˆžπ”Όπ›Όξ‚ƒξ€œπ‘‘π‘‘+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿπ›Ό,(𝑁)ξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–=ξ€œξ‚ξ‚„π‘‘π‘‘+𝑠limπ›Όβ†’βˆžπ”Όπ›Όξ‚ƒπ‘ξ‚€π‘Ÿ,π‘‹π‘Ÿπ›Ό,(𝑁)𝑙𝑖=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–=ξ€œξ‚ξ‚„π‘‘π‘Ÿπ‘‘π‘‘+π‘ π”Όξ‚ƒπ‘ξ‚€π‘Ÿ,π‘‹π‘Ÿ(𝑁)𝑙𝑖=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ƒξ€œξ‚ξ‚„π‘‘π‘Ÿ=𝔼𝑑𝑑+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿ(𝑁)ξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„(5.25)which along with (5.22) and (5.23) implieslimπ›Όβ†’βˆžπ”Όπ›Ό[𝑛𝛼(𝑑+𝑠)βˆ’1𝑖=𝑛𝛼(𝑑)𝑏𝛼𝑑𝛼𝑖,𝑋𝑑𝛼,(𝑁)π›Όπ‘–ξ‚πœ–π›Όπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π›Όπ‘‘π‘–ξ‚ξ‚ƒξ€œ]=𝔼𝑑𝑑+π‘ π‘ξ‚€π‘Ÿ,π‘‹π‘Ÿ(𝑁)ξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„.(5.26)This, (5.20), and (5.21) ensure (5.19).Step 3. It remains to show (1.8). By the uniqueness of the Doob-Meyer decomposition, 𝑀 has the required compensator if and only if𝔼𝑀2𝑑+π‘ βˆ’π‘€2π‘‘βˆ’ξ€œπ‘‘π‘‘+π‘ π‘Ž2ξ‚€π‘Ÿ,π‘‹π‘Ÿξ‚ξ‚π‘‘π‘Ÿπ‘™ξ‘π‘–=1β„Žπ‘–ξ‚€π‘‹π‘‘π‘–ξ‚ξ‚„=0(5.27)holds for all 0≀𝑑1<⋯≀𝑑𝑙≀𝑑, 𝑠β‰₯0, 𝑙β‰₯1 and bounded β„Ž1,…,β„Žπ‘™βˆˆπΆ(ℝ). Now, the discrete analogue of (5.27) for 𝔼𝛼, π‘Žπ›Ό, and 𝑋𝛼 holds. Proceeding similarly to the proof of (5.11) one can show that the left-hand side of this equation converges to the left-hand side of (5.27) as π›Όβ†’βˆž. Therefore, we obtain (5.27). For the sake of brevity we omit the details. It should be mentioned, however, that we now need uniform integrability of {(π‘‹π›Όπ‘Ÿ)2βˆΆπ‘Ÿβ‰€π‘‘+𝑠,𝛼β‰₯1}. This is why we established (5.2) for π‘ž being strictly larger than 2.

The assumptions of Proposition 5.2 can be checked with help of the following two lemmas, where β„šπ›Ό and β„š refer to any laws on 𝐷(ℝ), and π‘Œπ›Ό and π‘Œ are the respective coordinate processes. By an abuse of notation, we denote the corresponding expectations by β„šπ›Ό and β„š either. The first lemma follows from [19, Theorem 3.8.8] and [19, Theorem 3.8.6(b) β‡’ (a)] along with Prohorov's theorem. Lemma 5.4 is more or less standard and can be proved with help of the continuity criterion 3.10.3 in [19]; we omit the details.

Lemma 5.3. Assume that (π‘Œπ›Όπ‘‘) is tight in ℝ for every rational 𝑑β‰₯0. Let π‘š>0, 𝛾>1, and assume for every 𝑇>0 that there is some finite constant 𝐢𝑇>0 such that for all 𝛼β‰₯1 and 𝑑,β„Žβ‰₯0 with 0β‰€π‘‘βˆ’β„Ž and 𝑑+β„Žβ‰€π‘‡, β„šπ›Όξ‚ƒ|||π‘Œπ›Όπ‘‘βˆ’β„Žβˆ’π‘Œπ›Όπ‘‘|||π‘š/2|||π‘Œπ›Όπ‘‘βˆ’π‘Œπ›Όπ‘‘+β„Ž|||π‘š/2ξ‚„β‰€πΆπ‘‡β„Žπ›Ύ.(5.28)Then (β„šπ›Ό) is tight.

Lemma 5.4. Let π‘š>0, 𝛾>1, and assume for every 𝑇>0 that there is some finite constant 𝐢𝑇>0 such that for all 𝛼β‰₯1 and 0β‰€π‘‘ξ…žβ‰€π‘‘ξ…žξ…žβ‰€π‘‡, limsupπ›Όβ†’βˆžβ„šπ›Όξ‚ƒ|||π‘Œπ›Όπ‘‘β€²βˆ’π‘Œπ›Όπ‘‘β€²β€²|||π‘šξ‚„β‰€πΆπ‘‡ξ‚€π‘‘ξ…žξ…žβˆ’π‘‘ξ…žξ‚π›Ύ.(5.29)Then if β„šπ›Όβ‡’β„š, the limit β„š has no mass outside of 𝐢(ℝ).

Proposition 5.5. (ℙ𝛼) is tight and each limit point has no mass outside of 𝐢(ℝ).

Proof. Let π‘ž>2 and 𝛿>1 satisfy (2.3). Using techniques as in the proof of Lemma 5.1 we can find a finite constant πΆπ‘ž>0 such that for every 0β‰€π‘‘ξ…žβ‰€π‘‘ξ…žξ…ž and 𝛼β‰₯𝛼0,𝔼𝛼|||π‘‹π›Όπ‘‘β€²βˆ’π‘‹π›Όπ‘‘β€²β€²|||π‘žξ‚„β‰€πΆπ‘ž{πœ–π›Όπ‘ž/2𝔼𝛼[(𝑛𝛼(𝑑′′)βˆ’1𝑖=𝑛𝛼(𝑑′)ξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||2)π‘ž/2]+πœ–π‘žπ›Όπ”Όπ›Ό[(𝑛𝛼(𝑑′′)βˆ’1𝑖=𝑛𝛼(𝑑′)ξ‚€|||𝑋1+𝛼𝑑𝛼𝑖|||)π‘ž]+𝑛𝛼(𝑑′′)βˆ’1𝑖=𝑛𝛼(𝑑′)ξ‚€1+𝔼𝛼|||𝑋𝛼𝑑𝛼𝑖|||π‘žπœ–ξ‚„ξ‚π›Ώβ‹€π‘žπ›Ό}.(5.30)Applying HΓΆlder's inequality to each of the first two summands on the right-hand side, using (5.2) and setting ⋀𝛿𝛾=(π‘ž/2), we may continue withβ‰€πΆπ‘ž{πœ–π›Όπ‘ž/2(𝑛𝛼(𝑑′′)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑′)𝔼𝛼|||𝑋1+𝛼𝑑𝛼𝑖|||2π‘ž/2ξ‚„)(𝑛𝛼(𝑑′′)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑′)1π‘ž/(π‘žβˆ’2))π‘ž/2βˆ’1+πœ–π‘žπ›Ό(𝑛𝛼(𝑑′′)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑′)𝔼𝛼|||𝑋1+𝛼𝑑𝛼𝑖|||ξ‚π‘žξ‚„)(𝑛𝛼(𝑑′′)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑′)1π‘ž/(π‘žβˆ’1))π‘žβˆ’1+𝑛𝛼(𝑑′′)βˆ’1βˆ‘π‘–=𝑛𝛼(𝑑′)(1+𝔼𝛼|||𝑋𝛼𝑑𝛼𝑖|||π‘žξ‚„)πœ–π›Ώβ‹€π‘žπ›Ό}β‰€πΆπ‘ž,π‘‡ξ‚†πœ–π›Όπ‘ž/2ξ‚€π‘›π›Όξ‚€π‘‘ξ…žξ…žξ‚βˆ’π‘›π›Όξ‚€π‘‘ξ…žξ‚ξ‚π‘ž/2+πœ–π‘žπ›Όξ‚€π‘›π›Όξ‚€π‘‘ξ…žξ…žξ‚βˆ’π‘›π›Όξ‚€π‘‘ξ…žξ‚ξ‚π‘ž+πœ–π›Ώβ‹€π‘žπ›Όξ‚€π‘›π›Όξ‚€π‘‘ξ…žξ…žξ‚βˆ’π‘›π›Όξ‚€π‘‘ξ…žβ‰€ξ‚ξ‚ξ‚‡πΆπ‘ž,π‘‡ξ‚€πœ–π›Όξ‚€π‘›π›Όξ‚€π‘‘ξ…žξ…žξ‚βˆ’π‘›π›Όξ‚€π‘‘ξ…žξ‚ξ‚ξ‚π›Ύβ‰€πΆπ‘ž,π‘‡π‘‘ξ‚€ξ‚€ξ…žξ…žβˆ’π‘‘ξ…žξ‚+πœ–π›Όξ‚π›Ύ,(5.31)where πΆπ‘ž,𝑇,πΆπ‘ž,𝑇>0 are some finite constants being independent of π‘‘ξ…ž,π‘‘ξ…žξ…žβ‰€π‘‡, and 𝛼β‰₯𝛼0. Then Lemma 5.4 ensures that any weak limit point of (ℙ𝛼) has no mass outside of 𝐢(ℝ). At this point, it is essential that we required π‘ž and 𝛿 to be strictly larger than 2, 1, respectively.
Toward the verification of tightness of (ℙ𝛼) we use HΓΆlder's inequality to get𝔼𝛼|||π‘‹π›Όπ‘‘βˆ’β„Žβˆ’π‘‹π›Όπ‘‘|||π‘ž/2|||π‘‹π›Όπ‘‘βˆ’π‘‹π›Όπ‘‘+β„Ž|||π‘ž/2≀𝔼𝛼|||π‘‹π›Όπ‘‘βˆ’β„Žβˆ’π‘‹π›Όπ‘‘|||π‘žξ‚„1/2𝔼𝛼|||π‘‹π›Όπ‘‘βˆ’π‘‹π›Όπ‘‘+β„Ž|||π‘žξ‚„1/2.(5.32)If β„Žβ‰₯πœ–π›Ό/2, then (5.31) implies that both factors on the right-hand side of (5.32) are bounded by πΆπ‘ž,𝑇(3β„Ž)𝛾/2. If β„Ž<πœ–π›Ό/2, then at least one of these factors vanishes since 𝑋𝛼 is constant on intervals of length πœ–π›Ό. Hence,𝔼𝛼|||π‘‹π›Όπ‘‘βˆ’β„Žβˆ’π‘‹π›Όπ‘‘|||π‘ž/2|||π‘‹π›Όπ‘‘βˆ’π‘‹π›Όπ‘‘+β„Ž|||π‘ž/2≀𝐢2π‘ž,𝑇3π›Ύξ‚β„Žπ›Ύ(5.33)for all 𝛼β‰₯𝛼0 and 𝑑,β„Žβ‰₯0 with 𝑑+β„Žβ‰€π‘‡. That is, (5.28) holds with π‘š=π‘ž. Therefore, Lemma 5.3 ensures tightness of (ℙ𝛼).

Appendix

Auxiliary Results

Here we give two auxiliary results. We first recall a square function inequality for martingales. Let 𝑀=(π‘€π‘›βˆΆπ‘›βˆˆβ„•0) be an (ℱ𝑛)-martingale on some probability space (Ξ©,β„±,β„™). The corresponding compensator is given by βŸ¨π‘€βŸ©π‘›=βˆ‘π‘›π‘–=1𝔼[(π‘€π‘–βˆ’π‘€π‘–βˆ’1)2βˆ£β„±π‘–βˆ’1].

Proposition A.1 (see [21, Theorem 2.11]). For every π‘ž>0 there is some finite constant πΆπ‘ž>0 depending only on π‘ž such that 𝔼max1≀𝑖≀𝑛|||𝑀𝑖|||π‘žξ‚„β‰€πΆπ‘žξ‚€π”Όπ‘€ξ‚­ξ‚ƒξ‚¬π‘›π‘ž/2+𝔼max1≀𝑖≀𝑛|||π‘€π‘–βˆ’π‘€π‘–βˆ’1|||π‘žξ‚„ξ‚.(A.1)

The second result is a Gronwall lemma for functions with discrete domain. It can be proven by means of iterating (A.2) 𝑛-times. We omit the proof since it is more or less well known.

Lemma A.2. Suppose 𝑔 is a mapping from β„•0 to ℝ+=[0,∞] with 𝑔(0)<∞. If there are finite constants 𝑐0,𝑐1β‰₯0 such that 𝑔(𝑛)≀𝑐0+𝑐1π‘›βˆ’1𝑖=0𝑔(𝑖)βˆ€π‘›β‰€π‘,(A.2)then 𝑔(𝑛)≀𝑐0ξ‚€1+𝑐1𝑛+𝑐𝑛1𝑔(0)<βˆžβˆ€π‘›β‰€π‘.(A.3)

Acknowledgment

The author thanks a referee for revealing a flaw of the original manuscript.