• No results found

Affine Markov processes on a general state space - 3: Markov processes

N/A
N/A
Protected

Academic year: 2021

Share "Affine Markov processes on a general state space - 3: Markov processes"

Copied!
19
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

Affine Markov processes on a general state space

Veerman, E.

Publication date

2011

Link to publication

Citation for published version (APA):

Veerman, E. (2011). Affine Markov processes on a general state space. Uitgeverij BOXPress.

General rights

It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open content license (like Creative Commons).

Disclaimer/Complaints regulations

If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You will be contacted as soon as possible.

(2)

Chapter

3

Markov processes

In this chapter we consider Markov processes and Feller processes. We develop some general theory for Markov processes where the state space assumes the par-ticular form X × Rp−m, as given in (3.2). This will be applied in Chapter 4 to

obtain the desired characterization for affine processes living on such a state space. After defining Markov transition functions, Markov semigroups and stochastic con-tinuity in Section 3.1, we derive some geometric properties of the state space in Section 3.2. Next we define in Section 3.3 the concept of regularity of a Markov process, similar to the definition of regularity for affine processes in [17, 38, 10], and we translate results of [38] from the affine setting to the more general setting. For this we make use of the symbol of a regular Markov process, a concept that can be found in [32]. We show in Proposition 3.14 that the symbol has a L´ evy-Khintchine representation under regularity conditions, which is used in Section 3.4 to characterize Feller processes as solutions of the martingale problem, see Theo-rem 3.20. This connects Chapter 2 with Chapter 3. For the proof of the theoTheo-rem, Proposition 2.5 will be crucial.

3.1

Definitions

As in the previous chapter, E ⊂ Rp is a closed set and E

∆ denotes the one-point

compactification.

Definition 3.1. A function pt(x, dz) on [0, ∞) × E × B(E) is called a

time-homogeneous Markov transition function on E if it satisfies the following prop-erties:

(3)

• pt(x, ·) is a substochastic measure on E (i.e. a measure with pt(x, E) ≤ 1),

for all t ≥ 0, x ∈ E;

• p0(x, ·) = δx, for x ∈ E;

• x 7→ pt(x, B) is Borel-measurable for B ∈ B(E), t ≥ 0;

• the Chapman-Kolmogorov relation holds, i.e. pt+s(x, B) =

Z

ps(y, B)pt(x, dy), t, s ≥ 0, x ∈ E, B ∈ B(E).

Likewise we define a Markov transition function on E∆, with the additional

re-quirement that pt(x, ·) is a probability measure on E∆.

A Markov transition function pt(x, dz) on E can always be extended to a

Markov transition function on E∆ by putting pt(∆, {∆}) = 1 and pt(x, ∆) =

1 − pt(x, E) for x ∈ E. One verifies that the above properties remain satisfied in

this extended setting. Conversely, if pt(x, dz) is a Markov transition function on

E∆such that pt(∆, {∆}) = 1 and pt(x, ∆) = 1 − pt(x, E) for x ∈ E, then pt(x, dz)

restricted to [0, ∞) × E × B(E) is a Markov transition function on E. With this in mind, we give the following definition of a time-homogeneous Markov process.

Definition 3.2. On some filtered measurable space (Ω, F , (Ft)t≥0), let (Px)x∈Ebe

a family of probability measures and let X be an E∆-valued adapted stochastic

pro-cess with Px(X0= x) = 1 for all x ∈ E. We call (X, (Px)x∈E) a time-homogeneous

Markov process with Markov transition function pt(x, dz) on E if

Px(Xt+s∈ B|FtX) = ps(Xt, B), Px-a.s., for x ∈ E, B ∈ B(E∆), (3.1)

where we put ps(∆, {∆}) = 1 and pt(x, ∆) = 1 − pt(x, E) for x ∈ E.

A Markov transition function pt(x, dz) on E induces a semi-group (Pt)t≥0

act-ing on functions f ∈ B(E) by

Ptf (x) =

Z

f (z)pt(x, dz), x ∈ E.

Indeed, the Chapman-Kolmogorov relation yields the semi-group property Pt+s =

PtPs. Note that we can write Ptf (x) = Exf (Xt) for f ∈ B(E), x ∈ E, in view

of (3.1), where (X, (Px)x∈E) is the corresponding Markov process. Here we put

f (∆) = 0 by convention. Hence if we let 1 denote the function in B(E) with constant value 1, then we can write

(4)

3.1. Definitions 53

If Pt1 = 1 for all t ≥ 0, then we call the Markov process (X, (Px)x∈E) and its

associated semi-group (Pt)t≥0 conservative.

It is worth noting that from Proposition 2.8 we infer that (X, (Px)x∈E) is

a Markov process when Px is the unique solution of the martingale problem for

(A, δx) in Ω given by (2.8) and X is the coordinate process. Indeed, define pt(x, dz)

by

pt(x, B) = Px(Xt∈ B), for t ≥ 0, x ∈ E∆, B ∈ B(E∆).

Then Proposition 2.8 yields measurability of x 7→ pt(x, B) and the Markov

prop-erty (3.1), which on its turn implies the Chapman-Kolmogorov relation. Since pt(∆, {∆}) = P∆(Xt= ∆) = 1, the restriction of pt(x, dz) to [0, ∞) × E × B(E) is

a Markov transition function on E, see the remark preceding Definition 3.2. Hence (X, (Px)x∈E) is a Markov process with transition function pt(x, dz) = Px(Xt∈ dz).

In this chapter we are interested in the converse situation, that is, we provide con-ditions such that a Markov process can be characterized as the solution of a certain martingale problem. This requires certain continuity properties of the process, to begin with stochastic continuity.

Definition 3.3. A Markov process (X, (Px)x∈E) is called stochastically continuous

if ps(x, ·) w

→ pt(x, ·) on E, for s → t, for all t ≥ 0, x ∈ E, that is, if

Z

f (z)ps(x, dz) →

Z

f (z)pt(x, dz),

as s → t, for all f ∈ Cb(E).

Remark 3.4. It is important to notice that there exist Markov processes with continuous sample paths in E∆, that are not stochastically continuous in the sense

of Definition 3.3. For example, consider the ODE

∂tf (t, x) = f (t, x)2, f (0, x) = x.

with x ∈ E := R+. Its unique solution is given by

f (t, x) =        0 for x = 0, t ≥ 0, (x−1− t)−1, for x > 0, 0 ≤ t < x−1, ∆ for x > 0, t ≥ x−1.

Take Ω = R+, F = B(R+) and define the filtration (Ft)t≥0 by Ft = B(R+) for

all t ≥ 0. Define probability measures (Px)x∈E on (Ω, F , (Ft)t≥0) by Px({x}) = 1

(5)

Px(Xt = f (t, x)) = 1 for all x ∈ E. Then (X, (Px)x∈E) is a Markov process on

E = R+ with Markov transition function pt(x, dz) on E given by

pt(x, B) =        1B(0) for x = 0, 1B((x−1− t)−1), for x > 0, 0 ≤ t < x−1, 0 for for x > 0, t ≥ x−1,

for B ∈ B(R+). The sample paths of X are continuous in E∆, Px-a.s. for all

x ∈ E. However, stochastic continuity does not hold. For example take x > 0, t = 1/x. Then R 1E(z)ps(x, dz) = 1 for all s < t, which does not converge to

R 1E(z)pt(x, dz) = 0 for s ↑ t.

3.2

Properties of the state space

Henceforth and throughout we assume that the state space E ⊂ Rp takes the

special form

E = X × Rp−m, (3.2) for some 1 ≤ m ≤ p, where X ⊂ Rmis a closed, convex set with non-empty interior

and such that

U := {u ∈ Cm: sup

x∈X

Re u>x < ∞} (3.3)

has non-empty interior. We decompose {1, . . . , p} = I ∪ J with I = {1, . . . , m}, J = {m+1, . . . , p} and accordingly a point x ∈ Cpis decomposed as x = (xI, xJ) ∈

Cm× Cp−m. For u ∈ Cp we let fu ∈ C(E) denote the function fu(x) = eu

>x

. Similarly we define fuI ∈ C(X ) and fuJ ∈ C(R

p−m), so that we have the

decom-position fuI,uJ(x) := fu(x) = fuI(xI)fuJ(xJ). Note that u ∈ U × iR

p−m if and

only if fu ∈ Cb(E). In particular it holds that iRm ⊂ U . On the other hand,

iRm∩ U= ∅. This is a consequence of the next lemma, which characterizes the

interior of U .

Lemma 3.5. It holds that fu∈ C0(X ) if and only if u ∈ U◦.

Proof. First we prove that fu∈ C0(X ) for u ∈ Re U◦. Fix u ∈ Re U◦. There exists

ε > 0 such that u ± εei∈ Re U◦ for all i. Therefore,

sup

x∈X

(6)

3.2. Properties of the state space 55

Let (xn) be a sequence in E such that |xn| → ∞. Arguing by contradiction,

suppose lim supn→∞u>xn> −∞. Then there exists a subsequence, also denoted

by (xn), such that lim infn→∞u>xn > −∞, and in addition xn,i → ∞ or xn,i→

−∞. Then (3.4) yields u>x

n → −∞, a contradiction. Hence u>xn → −∞

whenever u ∈ Re U◦ and |x

n| → ∞ with xn∈ E, in other words, fu ∈ C0(X ) for

all u ∈ Re U◦. For u ∈ U◦ we have |fu(x)| ≤ fRe u(x), so by the previous we have

fu∈ C0(X ) for all u ∈ U◦.

Now suppose u ∈ U \U◦. Translating X does not affect U , so we may assume without loss of generality that 0 ∈ X . There exists v ∈ B(0, 1) such that u + εv 6∈ U for all ε > 0. This yields the existence of a sequence xn ∈ X such that

(u + v/n)>xn= n. If u>xn6→ −∞, then we are done. If u>xn→ −∞, then there

exists N > 0 such that u>xn< −1 for all n ≥ N . Henceexn = −xn/(u

>x n) ∈ X ,

for n ≥ N , by convexity of X and the fact that 0 ∈ X and xn ∈ X . We have

|exn||v| ≥ v>xen= −n

2/(u>x

n) + n ≥ n,

soxen→ ∆. On the other hand, u>xen= −1, which yields that fu6∈ C0(X ), as we needed to show.

Corollary 3.6. If u ∈ U◦ and v ∈ U , then v + λu ∈ Ufor all λ ∈ (0, 1].

Proof. Translating X does not affect U , so we may assume without loss of gener-ality that 0 ∈ X . Let |xn| → ∞, with xn ∈ X and λ ∈ (0, 1] be arbitrary. Since

0 ∈ X , it holds that yn:= λxn∈ X and |yn| → ∞. Hence the result follows from

Lemma 3.5, as fv+λu(xn) = fv(xn)fu(yn) → 0 for n → ∞, since fv ∈ Cb(X ) and

fu∈ C0(X ).

As we saw in Lemma 3.5, there is equivalence between u ∈ U◦and fu∈ C0(X ).

The following lemma shows that a similar equivalence holds between convergence in U◦ and convergence in C0(X ).

Lemma 3.7. It holds that un → u in U◦ if and only if fun → fu in C0(X ) (i.e.

kfun− fuk∞→ 0).

Proof. The “if”-part is easy. We show the “only if”-part. First assume un and

u are real-valued. Fix ε > 0 arbitrarily. We first show that there exists M > 0 and N > 0 such that for all x ∈ X with |x| ≥ M it holds that |fu(x)| < ε and

|fun(x)| < ε for all n ≥ N . Let δ > 0 be such that vi := u + δei ∈ U

and

wi:= u − δei ∈ U◦, for all i = 1, . . . , p. By Lemma 3.5 we have fvi, fwi ∈ C0(X )

(7)

such that for all x ∈ X with |x| > M it holds that |fu(x)| < ε, |fvi(x)| < ε and

|fwi(x)| < ε, for all i, i.e.

u>x ≤ log ε, v>i x ≤ log ε, w>i x ≤ log ε, for all x ∈ X . (3.5)

For all n ∈ N we can write

un= u + p X i=1 λniei− p X i=1 µniei,

with λni, µni ≥ 0. Since un → u, there exists N > 0 such that both P p i=1λ n i and Pp i=1µ n

i are bounded by δ/2 for n ≥ N . Fix n ≥ N and define [0, 1]-valued

numbers ti= λni/δ, si= µni/δ, t0= 1 − p X i=1 (ti+ si).

Then we can write

un= t0u + p X i=1 tivi+ p X i=1 siwi,

so that it follows that for all x ∈ X with |x| > M we have u>

nx ≤ log ε in view of

(3.5). Hence |fun(x)| ≤ ε for |x| > M and n ≥ N .

Next we write

fun(x) − fu(x) = (e

(un−u)>x− 1)eu>x.

For x ∈ X with |x| > M , the left-hand side is bounded by 2ε for n ≥ N . For |x| ≤ M we can bound the right-hand side by

exp(|(un− u)>x|) exp(|u>x|) ≤ exp(M |un− u|) exp(M |u|),

which can be made arbitrarily small by choosing n large enough. It follows that fun(x) → fu(x) uniformly in x.

Now assume un and u are complex-valued. Let un= an+ ibn, u = a + ib, with

an→ a in Re U◦and bn→ b in Rp. We can write

fun(x) − fu(x) = (e

a>nx− ea>x)eib>nx+ e(a+ib)>x(ei(bn−b)>x− 1),

and bound it by

|ea>nx− ea>x| + C|b n− b|,

with C = exp(supx∈Xa>x) < ∞. Both terms tend to zero uniformly in x, which

(8)

3.2. Properties of the state space 57

Corollary 3.8. For all compact K ⊂ U◦ it holds that supu∈Kkfuk∞< ∞.

Proof. By Lemma 3.7 it holds that u 7→ kfuk∞ is continuous on U◦. This

imme-diately gives the result.

We now investigate continuity properties of u 7→ R fu(x)µ(dx), where µ is

some finite measure on E. An immediate consequence of Lemma 3.7 is that u 7→ R fu(x)µ(dx) is continuous on U◦× iRp−m. However, continuity on U × iRp−m

fails in general, as is demonstrated in the next example.

Example 3.9. Consider E = X = {x ∈ R2 : x

2 ≥ x21}. Then one verifies that

Re U = (R × (−∞, 0)) ∪ {(0, 0)}. Define the finite measure µ on E with support {(k, k2

) : k ∈ N} by

µ({(k, k2)}) = 1/k2, for k ∈ N.

Take un= (n−1, −n−4) ∈ Re U , which converges to u0= 0 ∈ Re U . Then it holds

that Z fu0(x)µ(dx) = ∞ X k=1 1 k2 = π2 6 , while Z fun(x)µ(dx) = ∞ X k=1 1 k2exp(k/n − k 2 /n4) ≥ 1 n4exp(n − 1) → ∞,

for n → ∞. HenceR fun(x)µ(dx) does not converge toR fu0(x)µ(dx).

Although continuity fails on the whole of U × iRp−m, it does hold on certain

compacta of U × iRp−m, as we show in the following lemma.

Lemma 3.10. Let u ∈ U , K ⊂ U◦ be compact and define the compact set C = {λu + (1 − λ)v : λ ∈ [0, 1], v ∈ K} ⊂ U . Then for all finite measures µ on E it holds that z 7→R fz(x)µ(dx) is continuous on C × iRp−m.

Proof. Let zn → z with zn, z ∈ C × iRp−m. Then zn,I = λnu + (1 − λn)vn, for

some λn∈ [0, 1], vn∈ K. It follows that

kfznk∞≤ kfuk∞kfvnk∞,

which is uniformly bounded in n ∈ N, in view of Corollary 3.8. Dominated con-vergence yields

Z

fzn(x)µ(dx) →

Z

fz(x)µ(dx),

(9)

The next continuity property will be needed later on.

Lemma 3.11. Let (X, (Px)x∈E) be a stochastically continuous Markov process

with state space E and let the compact set C ⊂ Rm be given as in Lemma 3.10.

Then for fixed x ∈ E, it holds that (t, u) 7→ Ptfu(x) is jointly continuous on

R+× C × iRp−m.

Proof. Let (tn, un) → (t, u) in R+× C × Rp−m. As in the proof of Lemma 3.10 it

holds that kfunk∞is uniformly bounded in n. We have to show that Ptnfun(x) →

Ptfu(x). By stochastic continuity we have ptn(x, dz)

w

→ pt(x, dz). Note that if

pt(x, E) = 0, then

|Ptnfun(x)| ≤ kfunk∞ptn(x, E) → 0 = Ptfu(x).

which yields the result. Therefore, suppose pt(x, E) > 0. Then we have ptn(x, E) >

0 for n large enough. Define probability measures on E by

e pn(dz) = ptn(x, dz) ptn(x, E) , ep(dz) =pt(x, dz) pt(x, E) . Then pen w

→p. By Skorohod’s Representation Theorem [34, Theorem 4.30] theree exist E-valued random variables Yn, Y defined on a common probability space

(Ω, F , P ) such that P ◦ Yn−1=pen, P ◦ Y−1=p and Ye n→ Y , P -a.s. By dominated convergence we have Z fun(z)epn(dz) = Z fun(Yn(ω))P (dω) → Z fu(Y (ω))P (dω) = Z fu(z)p(dz).e

This gives the result.

3.3

The symbol of a regular Markov process

In this section we consider regular Markov processes. Intuitively speaking, such a process behaves locally like a L´evy process. Recall that if (X, (Px)x∈E) is a L´evy

process with L´evy triplet (b, c, K) with respect to a truncation function χ, then its Fourier-Laplace transform can be written as

Ptfu(x) = fu(x) exp(tΨ(u)),

for u ∈ U × iRp−m, where Ψ is of the L´evy-Khintchine form

Ψ(u) = u>b +12u>cu + Z

(10)

3.3. The symbol of a regular Markov process 59

Hence, we call a Markov processes regular if this holds “locally”, that is, if

Ptfu(x) ≈ fu(x) exp(tΨ(u, x)),

for some Ψ(u, x), t ≥ 0 small enough. We make this precise in the next definition.

Definition 3.12. We call (X, (Px)x∈E) regular if the right-hand derivative

∂+t

t=0Ptfu(x)

exists for all x ∈ E, u ∈ U◦× iRp−m, and is continuous in u for all x ∈ E. In that

case we define the symbol of the process X by

Ψ(u, x) = ∂ + t t=0Ptfu(x) fu(x) , (3.6)

for u ∈ U◦× iRp−m, x ∈ E. We call X strongly regular if it is regular and in

addition if for all x ∈ E, y ∈ U◦, compact K ⊂ Rp−m, there exists T > 0 such that sup z∈K,t∈[0,T ] 1 t(Ptfy,iz(x) − fy,iz(x)) < ∞.

Note that our definition of regularity slightly differs from the definition in [17], as the parameter u is restricted to U◦× iRp−m. Moreover, one usually defines the

symbol with a minus-sign on the right-hand side of (3.6), see [32]. Similar as for L´evy processes, the symbol Ψ of a regular process is also of the L´evy-Khintchine form. This is proved in Proposition 3.14 below, for which we need the following lemma. The proofs are based on the proof of [38, Theorem 4.3]. Recall that µn

converges vaguely to µ, denoted by µn v

→ µ, if Z

f (z)µn(dz) →

Z

f (z)µ(dz), for all f ∈ C0(E).

Lemma 3.13. Let µn be a sequence of finite measures on E and suppose

Z

fu(z)µn(dz) → Φ(u),

for all u ∈ U◦× iRp−m, with Φ : U× iRp−m → C a continuous function. Then

there exists a finite measure µ such that

(i) fθ,0(z)µn(dz) w → fθ,0(z)µ(dz) for all θ ∈ Re U◦, (ii) µn v → µ,

(11)

(iv) If in addition µn(E) → µ(E), then µn w

→ µ.

Proof. Let θ ∈ Re U◦ be arbitrary. By assumption, the characteristic functions of the exponentially tilted measures fθ,0(z)µn(dz) converge to u 7→ Φ(θ + uI, uJ),

which is continuous in 0 as u 7→ Φ(u) is continuous on U◦× iRp−m. By L´evy’s

continuity theorem we have

fθ,0(z)µn(dz) w

→ µθ(dz) (3.7)

for some measure µθ(dz) with characteristic function u 7→ Φ(θ + uI, uJ). Let µ be

a vague accumulation point of µn, which exists by Helly’s selection theorem (see

[4, Corollary 5.7.6]). Then it holds that

fθ,0(z)µn(dz) v

→ fθ,0(z)µ(dz), (3.8)

along a subsequence. Since weak convergence implies vague convergence, we infer from the uniqueness of the vague limit that

µθ(dz) = fθ,0(z)µ(dz), (3.9)

by combining (3.7) and (3.8). This yields the first assertion. Moreover, we conclude that µnhas only one vague accumulation point µ, which yields the second assertion.

Since µθ(dz) has characteristic function u 7→ Φ(θ + uI, uJ), it follows from (3.9)

that Φ(θ + uI, uJ) = Z fu(z)µθ(dz) = Z fθ+uI,uJ(z)µ(dz),

for all u ∈ iRp. This gives the third assertion, as θ ∈ Re U◦was chosen arbitrarily. The last assertion is a consequence of [41, II.6.8] (but notice the difference in terminology).

Proposition 3.14. Let (X, (Px)x∈E) be a regular Markov process with symbol Ψ.

Then there exist functions b : E → Rp, c : E → Sp+, γ : E → R+, a transition

kernel K from E to Rp\{0} with supp K(x, dz) ⊂ E − x and

Z (|z|2∧ 1)K(x, dz) < ∞, (3.10) such that Ψ(u, x) = u>b(x) +12u>c(x)u − γ(x) + Z (eu>z− 1 − u>χ(z))K(x, dz), (3.11) for u ∈ U◦× iRp−m, x ∈ E, where χ denotes a continuous truncation function.

(12)

3.3. The symbol of a regular Markov process 61

Proof. Fix x ∈ E and let tn ↓ 0. We define measures Kn(x, dz) on Rp\{0} with

supp Kn(x, dz) ⊂ E − x, by Kn(x, B) = 1 tn Z 1B(z − x)ptn(x, dz), for B ∈ B(R p\{0}).

Note that (3.6) can be rewritten as

Ψ(u, x) = lim n→∞ Z (eu>z− 1)Kn(x, dz)) + 1 tn (ptn(x, E) − 1)  , for u ∈ U◦× iRp−m. Let e

µn(x, dz) be the compound Poisson distribution with

L´evy measure Kn(x, dz) and define the infinitely divisible substochastic measure

µn(x, dz) by µn(x, dz) = exp( 1 tn (ptn(x, E) − 1))µen(x, dz) Then we have exp(Ψ(u, x)) = lim n→∞exp Z (eu>z− 1)Kn(x, dz)  exp 1 tn (ptn(x, E) − 1)  = lim n→∞ Z fu(z)µn(x, dz),

for u ∈ U◦× iRp−m. By Lemma 3.13, there exists a finite measure µ(x, dz) that

is the vague limit of µn(x, dz). Moreover, for all θ ∈ Re U◦ we have

fθ,0(z)µn(x, dz) w

→ fθ,0(z)µ(x, dz).

Since the class of infinitely divisible measures is closed under exponentially tilting and weak convergence, it follows that fθ,0(z)µ(x, dz) is infinitely divisible and

hence the same holds for µ(x, dz). In addition, we have by Lemma 3.13 that

exp(Ψ(u, x)) = Z

fu(z)µ(x, dz), (3.12)

for u ∈ U◦× iRp−m. In particular µ(x, E) > 0. Since µ(x, E) is the vague limit

of a sequence of substochastic measures, we also have µ(x, E) ≤ 1. Therefore, γ(x) := − log µ(x, E) ≥ 0. By the L´evy-Khintchine formula, there exist b(x) ∈ Rp, c(x) ∈ S+p and a measure K(x, dz) on B(Rp\{0}) satisfying (3.10) such that

Z fu(z)µ(x, dz) = exp  − γ(x) + u>b(x) +12u>c(x)u + Z (eu>z− 1 − u>χ(z))K(x, dz)  ,

(13)

for u ∈ iRp

and even for u ∈ U × iRp−m in view of [48, Theorem 27.17]. Here χ

denotes a truncation function, which we choose to be continuous. Combining the above display with (3.12) gives

Ψ(u, x) + k(x, u)πi = γ(x) + u>b(x) +1 2u

>c(x)u +Z (eu>z

− 1 − u>χ(z))K(x, dz), for some k(x, u) ∈ 2Z, possibly depending on x and u. Since Ψ(u, x) as well as the right-hand side of the above display are continuous in u on U◦× iRp−m by

Lemma 3.7, it holds that k(x, u) does not depend on u. In addition, choosing u ∈ Re U◦× {0} gives k(x, u) = 0, as all other terms are real-valued. Thus we have derived that

Ψ(u, x) = γ(x) + u>b(x) +1 2u

>c(x)u +Z (eu>z

− 1 − u>χ(z))K(x, dz). Since χ is chosen to be continuous, [48, Theorem 8.7] yields

Z

f (z)Kn(x, dz) →

Z

f (z)K(x, dz), as n → ∞,

for f ∈ Cb(Rp) vanishing on a neighborhood of 0. Note that if f (z) = 0 for z ∈ E −

x, then the left-hand side of the above display is zero, since supp Kn(x, dz) ⊂ E −x.

This yields that the support of K(x, dz) is contained in E − x and we have proved the result.

3.4

Feller processes

In this section we recall the definition of a Feller process and provide conditions under which a (regular) Feller process solves a certain martingale problem.

Definition 3.15. A Markov process (X, (Px)x∈E) is called a Feller process and

the corresponding semigroup (Pt)t≥0is called a Feller semi-group if

(i) PtC0(E) ⊂ C0(E) for all t ≥ 0;

(ii) Ptf (x) → f (x) as t ↓ 0, uniformly in x ∈ E, for f ∈ C0(E).

With every Feller semi-group one can associate its infinitesimal generator , which is the linear operator A on C0(E) (i.e. linear operator from a subspace D(A) ⊂ C0(E)

to C0(E)) defined by

Af = lim

t↓0

Ptf − f

(14)

3.4. Feller processes 63

where the limit is taken with respect to the supremum norm. The domain D(A) ⊂ C0(E) of A is the subspace of those f ∈ C0(E) for which this limit exists.

Note that for a Feller process (X, (Px)x∈E) we may take Ω to be of the form

(2.8) and X the coordinate process, in view of [34, Theorem 19.15] and the ensuing paragraph. In that case we call (X, (Px)x∈E) a canonical Feller process. For a

stochastically continuous Markov process to be Feller, it suffices to have Ptf ∈

C0(E) for f in a certain subclass of C0(E). This is the content of Proposition 3.17

below, the proof of which uses the following lemma, taken from [38, Theorem 3.5].

Lemma 3.16. Define the class of functions H by

H = {x 7→ fy(xI)

Z

fiz(xJ)g(z)dz : y ∈ U◦, g ∈ Cc∞(R

p−m)}. (3.14)

It holds that the linear span of H is dense in C0(E) with respect to the supremum

norm.

Proof. Lemma 3.5 yields that xI 7→ fy(xI) ∈ C0(X ) for all u ∈ U◦, while xJ 7→

R fiz(xJ)g(z)dz ∈ C0(Rp−m) for all g ∈ Cc∞(Rp−m) by the Riemann-Lebesgue

Lemma. Hence H ⊂ C0(E). Note that the functions in H can be written as

x 7→ fy(xI)bg(xJ), where bg denotes the Fourier transform of g. Therefore, the product of two elements

x 7→ fy1(xI)gb1(xJ) and x 7→ fy2(xI)gb2(xJ) can be written as

x 7→ fy1+y2(xI) \g1∗ g2(xJ),

where g1 ∗ g2 denotes the convolution. Since U◦ is closed under addition and

Cc(Rp−m) is closed under convolution, it follows that H is closed under multipli-cation, i.e. H is a subalgebra of C0(E). In addition, H vanishes nowhere (i.e. there

is no x0 ∈ E such that h(x0) = 0 for all h ∈ H) and H separates points (i.e. for

all x, y ∈ E with x 6= y there exists h ∈ H with h(x) 6= h(y)). Since E is closed, it is locally compact, whence the result follows from a locally compact version of the Stone-Weierstrass Theorem (see [49, Corollary 7.3.9]).

Proposition 3.17. Let (X, (Px)x∈E) be a Markov process and let H be given by

(3.14). Then X is a Feller process if

(i) Pt(H) ⊂ C0(E) for all t ≥ 0;

(15)

Proof. We check the properties stated in Definition 3.15. Let f ∈ C0(E).

Accord-ing to Lemma 3.16 there exists a sequence fn in the linear span of H such that

kfn− f k∞ → 0, as n → ∞. By the contraction property of the semigroup Pt

we have kPtfn − Ptf k∞ ≤ kfn− f k → 0. Since Ptfn ∈ C0(E) and C0(E) is a

complete space with respect to the supremum norm, it follows that Ptf ∈ C0(E),

which gives the first property. Stochastic continuity immediately yields

Ptf (x) =

Z

f (z)pt(x, dz) →

Z

f (z)p0(x, dz) = f (x), for f ∈ Cb(E),

so in particular for f ∈ C0(E). The uniform convergence follows from [34,

Theo-rem 19.6], whence we have shown the second property.

In Theorem 3.20 below we provide conditions such that the generator of a regular Feller process (X, (Px)x∈E) assumes the form (2.2) for f ∈ Cc2(E) and we

show that Px is a solution of the corresponding martingale problem. The special

form of the generator is a consequence of the L´evy-Khintchine form of the symbol Ψ, while we use Proposition 2.5 to deduce that Px solves the martingale problem.

Lemma 3.18. Let f ∈ Ck,0

(Rp×Rq

) and define g : Rp→ C by g(x) =R

Cf (x, y)dy,

for some compact set C ⊂ Rq. Then it holds that g ∈ Ck

(Rp) and

∂xg(x) =

Z

C

∂xf (x, y)dy.

Proof. It is sufficient to prove this for k = p = 1. By Fubini we can write

g(x) = g(0) + Z C Z x 0 ∂tf (t, y)dtdy = g(0) + Z x 0 Z C ∂tf (t, y)dydt,

since (t, y) 7→ ∂tf (t, y) is continuous, whence bounded on [0, x] × C. Dominated

convergence yields that t 7→RC∂tf (t, y)dy is continuous, so that we infer from the

fundamental theorem of calculus that g is continuously differentiable with

∂xg(x) =

Z

C

∂xf (x, y)dy.

Lemma 3.19. Let K be a transition kernel from E to Rp\{0} with supp K(x, dz) ⊂

E − x, and satisfying (3.10). Let f ∈ Cb2,0(E × Rq

) and define g : E → C by g(x) =R

Cf (x, y)dy, for some compact set C ⊂ R

q. Then it holds that g ∈ C2(E)

and Z (g(x + z) − g(x) − ∇g(x)>χ(z))K(x, dz) = Z C Z (f (x + z, y) − f (x, y) − ∇f (x, y)>χ(z))K(x, dz)dy.

(16)

3.4. Feller processes 65

Proof. By Lemma 3.18 we have g ∈ C2(E) and ∇g(x) =R

C∇f (x, y)dy. Therefore, Z (g(x + z) − g(x) − ∇g(x)>χ(z))K(x, dz) = Z Z C (f (x + z, y) − f (x, y) − ∇f (x, y)>χ(z))dyK(x, dz).

It remains to show that the order of integration can be interchanged by Fubini’s theorem. Fix x0∈ E. SinceR (|z|2∧1)K(x0, dz) < ∞ and supp K(x0, dz) ⊂ E−x0,

it suffices to show that

|f (x0+ z, y) − f (x0, y) − ∇f (x0, y)>χ(z)| ≤ M (|z|2∧ 1), (3.15)

for some M > 0, for all z ∈ E − x0, y ∈ C. Recall that χ(z) = z in a neighborhood

of 0, so there exists ε > 0 such that we can write

f (x0+ z, y) − f (x0, y) − ∇f (x0, y)>χ(z) = X i,j Z 1 0 Z 1 0 fij(x0+ stz, y)tzizjds dt,

for all |z| ≤ ε, where fij is short-hand notation for ∂xi∂xjf . By assumption

(x, y) 7→ fij(x, y) is continuous. In particular it is bounded on B(x0, ε) × C.

It follows that

|f (x0+ z, y) − f (x0, y) − ∇f (x0, y)>χ(z)| ≤ M1|z|2,

for some M1 > 0, for all |z| ≤ ε, y ∈ C. Since f and χ are bounded and since

y 7→ ∇f (x0, y) is bounded on C by continuity, we have

|f (x0+ z, y) − f (x0, y) − ∇f (x0, y)>χ(z)| ≤ M2,

for some M2 > 0, for all |z| > ε, y ∈ C. This yields (3.15) and the lemma is

proved.

Theorem 3.20. Let (X, (Px)x∈E) be a strongly regular canonical Feller process,

with infinitesimal generator A and symbol Ψ given by (3.11), for measurable func-tions b : E → Rp, c : E → Sp

+, γ : E → R+, a transition kernel K from E

to Rp\{0} with supp K(x, dz) ⊂ E − x, and satisfying (2.1). In addition, let A]: C2

b(E) → M (E) be the linear operator given by

A]f (x) = ∇f (x)>b(x) +1 2tr (∇ 2f (x)c(x)) − γ(x)f (x) + Z (f (x + z) − f (x) − ∇f (x)>χ(z))K(x, dz).

Assume A](Cc2(E)) ⊂ C0(E) as well as A](H) ⊂ C0(E), where H is given by

(3.14). Then C2

c(E) ⊂ D(A) and A|C2

c(E) = A

]| C2

c(E). In particular, Px is a

(17)

Proof. Throughout we make use of the equality Z A]f y,iz(x)g(z)dz = A] Z fy,iz(x)g(z)dz  , for y ∈ U◦, g ∈ Cc(Rp−m), (3.16)

which is a consequence of Lemmas 3.18 and 3.19. We first show that H ⊂ D(A) and that

Ah = A]h, for all h ∈ H. (3.17)

Let y ∈ U◦, g ∈ Cc(Rp−m) and take h(x) =R fy,iz(x)g(z)dz. Then h ∈ C0(E) by

Lemma 3.16. Note that for u ∈ U × iRp−m we can write A]fu(x) = fu(x)  u>b(x) +1 2u >c(x)u − γ(x) + Z (eu>z− 1 − u>χ(z))K(x, dz)  . (3.18)

Fubini and strong regularity of X yields 1

t(Pth(x) − h(x)) = Z 1

t(Ptfy,iz(x) − fy,iz(x))g(z)dz → Z

Ψ(y, iz, x)fy,iz(x)g(z)dz,

as t ↓ 0, for all x ∈ E. From (3.11) we infer that Ψ(u, x)fu(x) = A]fu(x) for

u ∈ U◦× iRp−m, x ∈ E. Together with (3.16) this gives

1

t(Pth(x) − h(x)) → A

]h(x),

as t ↓ 0, for all x ∈ E. Since (Pt)t≥0 is a Feller semi-group with generator A

and A]h ∈ C0(E) by assumption, the pointwise convergence in the above display

suffices for uniform convergence, see [48, Lemma 31.7]. Hence h ∈ D(A) and (3.17) holds.

Let Tn be the stopping time given by (2.7), n ∈ N. By Dynkin’s formula [34,

Lemma 19.21], it holds for all h ∈ H that

h(XTn

t ) − h(X0) −

Z t∧Tn

0

Ah(Xs)ds

is a Px-martingale for all x ∈ E. Let (y, iz0) ∈ U◦× iRp−m be arbitrary. Take

a sequence of non-negative functions (gk) in Cc∞(Rp−m) with support contained

in the unit ball B(0, 1) in Rp−m, withR gk(z)dz = 1 for all k ∈ N and such that

gk(z)dz converges weakly to δz0(dz). We define a sequence (hk) in H by

hk(x) =

Z

(18)

3.4. Feller processes 67 We show that hk(XtTn) − Z t∧Tn 0 A]h k(Xs)ds L1 → fy,iz0(X Tn t ) − Z t∧Tn 0 A]f y,iz0(Xs)ds, (3.19)

as k → ∞, so that the right-hand side, being an L1

-limit of a sequence of Px

-martingales, is a Px-martingale itself for all x ∈ E. Since hk(x) → fy,iz0(x) and

khkk∞= kfyk∞< ∞, we have hk(XtTn) L1 → fy,iz0(X Tn t ) as k → ∞. By (3.16) it holds that Z t∧Tn 0 A]h k(Xs)ds = Z t∧Tn 0 Z A]f y,iz(Xs)gk(z)dzds.

From (3.18) together with (2.1) it follows that z 7→ A]fy,iz is continuous, whence

bounded on B(0, 1). Hence

Z

A]fy,iz(x)gk(z)dz → A]fy,iz0(x),

as gk(z)dz w

→ δz0(dz). It also follows from (3.18) together with (2.1) that

sup |x|≤n,|z|≤1 |A]f y,iz(x)| := C < ∞. This gives Z A]f y,iz(Xs)gk(z)dz ≤ C Z gk(z)dz = C,

for s < Tn. From dominated convergence we deduce that

Z t∧Tn 0 Z A]f y,iz(Xs)gk(z)dzds L1 → Z t∧Tn 0 A]f y,iz0(Xs)ds,

whence we have that (3.19) holds. It remains to show that fu(XtTn) −

Rt∧Tn

0 A

]f

u(Xs)ds is a Px-martingale for

all x ∈ E, u ∈ iRp, n ∈ N. Indeed, in that case, Proposition 2.5 yields that Px is

a solution of the martingale problem for (A], δ

x) for all x ∈ E, while Lemma 2.11

gives A]f (x) = ∂t+ t=0Exf (Xt) = ∂ + t t=0Ptf (x), for f ∈ C 2 c(E).

Since (Pt)t≥0 is a Feller semi-group with generator A and A](C0(E)) ⊂ C0(E)

(19)

uniform convergence, see [48, Lemma 31.7], whence f ∈ D(A) and A]f = Af for

f ∈ C2 c(E).

Let u0∈ iRpbe arbitrary and take a sequence (uk) in U◦×iRp−mthat converges

to u0 and with the property that

sup

k∈N

kfukk∞< ∞, (3.20)

which is possible in view of the proof of Lemma 3.10. Since χ(z) = z in a neigh-borhood of 0, there exists ε > 0 such that for |z| ≤ ε it holds that

|eu>kz− 1 − u> kχ(z)| ≤ |uk|2|z|2 Z 1 0 Z 1 0 exp(u>kzst)t ds dt ≤ C|z|2,

for some constant C > 0, uniformly in k. Also, since χ is bounded and since we have (3.20), there exists C > 0 such that for all x ∈ E , |z| > ε with z ∈ E − x, k ∈ N, we have that |fuk(x)(e u>kz− 1 − u> kχ(z))| = |fuk(x + z) − fuk(x)(1 + u > kχ(z))| ≤ C.

By assumption supp K(x, dz) ⊂ E − x, whence by the previous we have

|fuk(x) Z (eu>kz− 1 − u> kχ(z))K(x, dz)| ≤ C Z (|z|2∧ 1)K(x, dz), for some C > 0, for all x ∈ E. Hence (3.18) together with (2.1) gives

sup

|x|≤n,k∈N

A]f

uk(x) < ∞,

while dominated convergence gives

Z (eu>kz− 1 − u> kχ(z))K(x, dz) → Z (eu>0z− 1 − u> 0χ(z))K(x, dz), as k → ∞. Hence A]fuk(x) → A ]f

u0(x) for all x ∈ E, as k → ∞, in view of

(3.18). Using dominated convergence again, we obtain

fuk(X Tn t ) − Z t∧Tn 0 A]f uk(Xs)ds L1 → fu(XtTn) − Z t∧Tn 0 A]f u(Xs)ds,

whence the right-hand side, being an L1

-limit of a sequence of Px-martingales, is

Referenties

GERELATEERDE DOCUMENTEN

Some of the biases the researcher brought to the research project included: student perceptions of summer learning, possible summer programs and resources parents would support,

integrative approach combining quantitative genetics, expression and diversity data permitted development of comprehensive gene networks for two major breeding traits, flowering

• Criminalization of poverty and focus on public safety that ignore the lack of safety for those living alone on the streets obscures key issues related to the social determinants

5 shows an example of super-transmission for three 40 nm slits separated by 40 nm in a 140 nm thick silver film, and the associated single slit transmission for Ag in the

Figure 4.15The required power to drive the deflecting voltage as a function of the input coupler quality factor and different beam loading

2B, although signals for neurine, choline, and phos- phocholine were observed, severe interferences from the matrix and other background ions were also seen in the lower mass

While other laboratory studies using similar or the same experimental task reported performance effects w ith small amounts of incentive money (Frisch &amp; Dickinson,

Let students know about the new citation management program that will help make writing bibliographies a