for branching symmetric Hunt processes with measure-valued branching rates

Zhen-Qing Chen∗, Yan-Xia Ren† and Ting Yang ‡§

Abstract

We establish weak and strong law of large numbers for a class of branching symmetric Hunt processes with the branching rate being a smooth measure with respect to the underlying Hunt process, and the branching mechanism being general and state-dependent. Our work is motivated by recent work on strong law of large numbers for branching symmetric Markov processes by Chen-Shiozawa [6] and for branching diffusions by Engl¨ander-Harris-Kyprianou [11]. Our results can be applied to some interesting examples that are covered by neither of these papers.

AMS 2010 Mathematics Subject Classification. Primary 60J25, Secondary 60J80.

Keywords and Phrases. Law of large numbers, branching Hunt processes, spine approach, h-transform, spectral gap

1 Introduction

1.1 Motivation

The law of large numbers (LLN) has been the object of interest for measure-valued Markov processes including branching Markov processes and . For branching Markov processes, the earliest work in this field dates back to 1970s when Watanabe [25, 26] studied the asymptotic properties of a branching symmetric diffusion, using a suitable Fourier analysis. Later, Asmussen and Hering [3] established an almost-sure limit theorem for a general supercritical under some regularity conditions. Recently, there is a revived interest in this field using modern techniques such as Dirichlet form method, martingales and spine method. Chen and Shiozawa [6] (see also [23]) used a Dirichlet form and spectral theory approach to obtain strong law of large numbers (SLLN) for branching symmetric Hunt processes. Among other assumptions, a spectral

∗Department of Mathematics, University of Washington, Seattle, WA 98195, USA. E-mail: [email protected] †LMAM School of Mathematical Sciences & Center for Statistical Science, Peking University, Beijing, 100871, P.R. China. E-mail: [email protected] ‡Corresponding author. School of Mathematics and , Beijing Institute of Technology, Beijing, 102488, P.R.China. Email: [email protected] §Beijing Key Laboratory on MCAACI, Beijing, 102488, P.R. China.

1 gap condition was used to obtain a Poincar´einequality that plays an important role in the proof of SLLN along lattice times. They proved SLLN holds for branching processes under the assumptions that the branching rate is given by a measure in Kato-class K (X) and the branching mechanism ∞ has bounded second moment. The spine method developed recently for measure-valued Markov processes is a powerful prob- abilistic tool in studying various properties of these processes; see, e.g., [9, 10, 11, 12, 14, 18, 19]. In [11], Engl¨ander, Harris and Kyprianou used spine method to obtain SLLN for branching (possi- bly non-symmetric) diffusions corresponding to the operator Lu + β(u2 u) on a domain D Rd − ⊂ (where β 0 is non-trivial) under certain spectral conditions. They imposed a condition on how far ≥ particles can spread in space (see condition (iii) on page 282 of [11]). That the underlying process is a diffusion plays an important role in their argument and the branching rate there has to be a function rather than a measure. The approach of [11] also involves p-th moment calculation with p > 1 which may not be valid for general branching mechanisms. Recently, Eckhoff, Kyprianou and Winkel [9] discussed the strong law of large numbers (SLLN) along lattice times for branching diffusions, which serves as the backbone or skeleton for superdiffusions. It is proved in [9, Theo- rem 2.14] that, if the branching mechanism satisfies a p-th moment condition with p (1, 2], the ∈ underlying diffusion and the support of the branching diffusion satisfy conditions similar to that presented in [11], then SLLN along lattice times holds. In this paper, we combine the functional analytic methods used in [6] with spine techniques to study weak and strong laws of large numbers for branching symmetric Hunt processes as well as the L log L criteria. This approach allows us to obtain new results for a large class of branching Markov processes, for which (i) the underlying spatial motions are general symmetric Hunt processes, which can be discontinuous and may not be intrinsically ultracontractive; (ii) the branching rates are given by general smooth measures rather than functions or Kato class measures; (iii) the offspring distributions are only assumed to have bounded first moments with no assumption on their second moments. In addition, we use L1-approach instead of Lp-approach for p (1, 2]. Now we describe ∈ the setting and main results of this paper in detail, followed by several examples illuminating the main results.

1.2 Branching symmetric Hunt processes and assumptions

Suppose we are given three initial ingredients: a Hunt process, a smooth measure and a branching mechanism. We introduce them one by one:

A Hunt process X: Suppose E is a locally compact separable metric space and E := E ∂ • ∂ ∪ { } is its one point compactification. m is a positive Radon measure on E with full support. Let X = (Ω, , ,θ , X , Π ,ζ) be a m-symmetric Hunt process on E. Here : t 0 H Ht t t x {Ht ≥ } is the minimal admissible filtration, θ : t 0 the time-shift operator of X satisfying { t ≥ } X θ = X for s,t 0, and ζ := inf t> 0 : X = ∂ the life time of X. Suppose for each t ◦ s t+s ≥ { t } t > 0, X has a symmetric transition density function p(t,x,y) with respect to the measure

2 m. Let P : t 0 be the Markovian transition semigroup of X, i.e., { t ≥ }

Ptf(x) := Πx [f(Xt)] = f(y)p(t,x,y)m(dy) ZE for any non-negative measurable function f. The symmetric Dirichlet form on L2(E,m) generated by X will be denoted as ( , ): E F

2 1 = u L (E,m) : lim (u(x) Ptu(x)) u(x)m(dx) < + , F ∈ t 0 t E − ∞ n → Z o 1 (u, v) = lim (u(x) Ptu(x)) v(x)m(dx), u, v . E t 0 t − ∈ F → ZE It is known (cf. [5]) that ( , ) is quasi-regular and hence is quasi-homeomorphic to a regular E F Dirichlet form on a locally compact separable metric space. In the sequel, we assume that X is m-irreducible in the sense that if A (E) has positive m-measure, then Π (T < + ) > 0 ∈B x A ∞ for all x E, where T := inf t> 0 : X A is the first hitting time of A. ∈ A { t ∈ } A branching rate µ: Suppose µ is a positive smooth Radon measure on (E, (E)). It uniquely • µ B determines a positive continuous additive functional (PCAF) At by the following Revuz formula: t 1 µ + f(x)µ(dx) = lim Πm f(Xs)dAs , f (E). t 0 t ∈B ZE → Z0  Here Π ( ) := Π ( )m(dx). m · E x · Offspring distributionsR p (x) : n 0 ,x E : Suppose p (x) : n 0 ,x E is a • {{ n ≥ } ∈ } {{ n ≥ } ∈ } family of mass functions such that 0 p (x) 1 and ∞ p (x) = 1. For ≤ n ≤ n=0 n each x E, p (x) : n 0 serves as the offspring distribution of a particle located at x. ∈ { n ≥ } P Let A(x) : x E be a collection of random variables taking values in 0, 1, 2, and { ∈ } { ···} distributed as P (A(x)= n)= pn(x).

Define + ∞ Q(x) := np (x), x E. (1.1) n ∈ n=0 X Throughout this paper we assume that the offspring distribution p (x) : n 0 satisfies the { n ≥ } following condition:

p0(x) 0, p1(x) 1 and sup Q(x) < . (1.2) ≡ 6≡ x E ∞ ∈ From these ingredients we can build a branching Markov process according to the following recipe: under a probability measure P , a particle starts from x E and moves around in E like x ∈ ∂ a copy of X. We use to denote the original particle, X (t) its position at time t and ζ its fission ∅ ∅ ∅ time. We say that splits at the rate µ in the sense that ∅ µ Px (ζ >t X (s) : s 0) = exp( At ). ∅ | ∅ ≥ −

3 When ζ ζ, it dies at time ζ. On the other hand, when ζ < ζ, it splits into a random ∅ ≥ ∅ number of children, the number being distributed as a copy of A(X (ζ )). These children, starting ∅ ∅− from their point of creation, will move and reproduce independently in the same way as their parents. If a particle u is alive at time t, we refer to its location in E as Xu(t). Therefore the time-t configuration is a E-valued X = X (t) : u , where is the set t { u ∈ Zt} Zt of particles alive at time t. With abuse of notation, we can also regard Xt as a random point X X measure on E defined by t := u t δXu(t). Let ( t)t 0 be the natural filtration of and ∈Z F ≥ = σ t : t 0 . Hence it defines a branching symmetric Hunt process X = (Ω, , t, Xt, Px) F∞ {F ≥ } P F∞ F on E with the motion component X, the branching rate measure µ and the branching mechanism function pn(x) : n 0 . When the branching rate measure µ is absolutely continuous with respect { ≥ } µ to m, i.e., µ(dy) = β(y)m(dy) for some non-negative function β, the corresponding PCAF At is t equal to 0 β(Xs)ds, and given that a particle u is alive at t, its probability of splitting in (t,t + dt) is β(Xu(tR))dt + o(dt). Since the function β determines the rate at which every particle splits, β is called the branching rate function in literature. Throughout this paper we use (E) (respectively, +(E)) to denote the space of bounded Bb B (respectively, non-negative) measurable functions on (E, (E)). Any function f on E will be B automatically extended to E by setting f(∂) = 0. We use f, g to denote f(x)g(x)m(dx) ∂ h i E and “:=” as a way of definition. For a, b R, a b := min a, b , a b := max a, b , and ∈ ∧ { } ∨ R { } log+ a := log(a 1). ∨ For every f +(E) and t 0, define ∈B ≥

Xt(f) := f(Xu(t)). u t X∈Z (Q 1)µ We define the Feynman-Kac semigroup Pt − by

(Q 1)µ (Q 1)µ + P − f(x) := Π exp A − f(X ) , f (E). (1.3) t x { t } t ∈B h i (Q 1)µ Since X has a transition density function, it follows that for each t> 0, Pt − admits an integral (Q 1)µ kernel with respect to the measure m. We denote this kernel by p − (t,x,y). The semigroup (Q 1)µ P − associates with a quadratic form ( (Q 1)µ, µ), where µ = L2(E; µ) and t E − F F F ∩

(Q 1)µ 2 µ − (u, u)= (u, u) u(x) (Q(x) 1)µ(dx), u . E E − − ∈ F ZE We say that a signed smooth measure ν belongs to the Kato class K(X), if

ν lim sup Πx At| | = 0. (1.4) t 0 x E ↓ ∈   For every non-negative ν K(X), we have Gαν < for every α > 0, where Gα is the ∈ k k∞ ∞ α-resolvent of X and G ν is the α-potential of ν. Define (u, u) := (u, u)+ α u2dm. By [24], α Eα E E R 2 u(x) ν(dx) Gαν α(u, u), u . (1.5) ≤ k k∞E ∈ F ZE 4 Thus when µ is in K(X), µ = , and the quadratic form ( (Q 1)µ, µ) is bounded from below. F F E − F For an arbitrary smooth measure µ, we define

(Q 1)µ µ 2 λ := inf − (u, u) : u with u(x) m(dx) = 1 . (1.6) 1 E ∈ F  ZE  Assumption 1. Let µ be a non-negative smooth measure on E so that the Schr¨odinger semigroup (Q 1)µ (Q 1)µ Pt − admits a symmetric kernel p − (t,x,y) with respect to the measure m and is jointly continuous in (x,y) E E for every t > 0. Moreover, λ ( , 0) and there is a positive ∈ × 1 ∈ −∞ continuous function h µ with h(x)2m(dx) = 1 so that (Q 1)µ(h, h)= λ . ∈ F E E − 1 Observe that if u is a minimizerR for (1.6), then so is u . Assumption 1 says that there is a | | minimizer for (1.6) that can be chosen to be positive and continuous. Clearly the following property holds for h: (h, v)= h(x)v(x)(Q(x) 1)µ(dx)+ λ h, v for every v µ. (1.7) E − 1h i ∈ F ZE (Q 1)µ µ The finiteness of λ1 implies that the bilinear form ( − , ) is bounded from below, and (Q 1)µ E F hence by [2], P − : t 0 is a strongly continuous semigroup on L2(E,m). Let σ( (Q 1)µ) { t ≥ } E − denote the spectrum of the self-adjoint operator associated with (Q 1)µ. Let λ be the second E − 2 bottom of σ( (Q 1)µ), that is, E − (Q 1)µ µ 2 λ2 := inf − (u, u) : u , u(x)h(x)m(dx) = 0, u(x) m(dx) = 1 . E ∈ F E E n Z Z o Assumption 2. There is a positive spectral gap in σ( (Q 1)µ): λ := λ λ > 0. E − h 2 − 1 (Q 1)µ Define h-transformed semigroup P h; t 0 from P − ; t 0 by { t ≥ } { t ≥ } λ1t h e (Q 1)µ + P f(x)= P − (hf)(x) for x E and f (E). (1.8) t h(x) t ∈ ∈B h 2 Then it is easy to see that Pt ; t 0 is an m-symmetric semigroup, where m := h m, and 1 is h { ≥ } an eigenfunction of Pt with eigenvalue 1. Furthermore the spectrum of the infinitesimal generator h 2 e e (Q 1)µ of Pt ; t 0 in L (E; m) is the spectrum of the infinitesimal generator of Pt − ; t 0 in 2 { ≥ } { ≥ } L (E; m) shifted by λ1. Hence under Assumption 2, we have the following Poincar´einequality: e h λht P ϕ 2 e e− ϕ 2 e (1.9) k t kL (E,m) ≤ k kL (E,m) for all ϕ L2(E, m) with ϕ(x)m(dx) = 0. ∈ E Remark 1.1. If the underlyingR process X satisfies that for each t > 0, the transition density e e function p(t,x,y) is bounded and is continuous in x for every fixed y E and that the branching ∈ rate measure µ is in the Kato class K(X) of X, then it follows from [1] that the Feymann-Kac (Q 1)µ p semigroup Pt − maps bounded functions to continuous functions and is bounded from L (E,m) to Lq(E,m) for any 1 p q + . By Friedrichs theorem, Assumptions 1 and 2 hold if we ≤ ≤ ≤ ∞ assume in addition that the embedding of ( , ) into L2(E,µ) is compact. F E1 Such an assumption is imposed in [6] to ensure the spectral gap condition and to obtain Poincar´e inequality (1.9).

5 1.3 Main results

Recall that ( t)t 0 is the natural filtration of X. Observe that (cf. [23, Lemma 3.3]) for every F ≥ x E and every f +(E), ∈ ∈B (Q 1)µ Ex [Xt(f)] = Pt − f(x). (1.10)

λ1t It is easy to see that Mt := e Xt(h) is a positive Px-martingale with respect to t. Let M := F ∞ limt + Mt. It is natural to ask when M is non-degenerate, that is, when Px(M > 0) > 0 for → ∞ ∞ ∞ (Q 1)µ x E? Under the assumptions that (i) m(E) < , (ii) the Feymann-Kac semigroup P − is ∈ ∞ t intrinsically ultracontractive, and (iii) h is bounded, it is proved in [19] that the condition

+ ∞ 2 + kpk(y)h(y) log (kh(y))µ(dy) < + (1.11) E E ∞ Z Z Xk=0 is necessary and sufficient for M to be non-degenerate. Condition (1.11) is usually called the ∞ L log L criteria. The first main result of this paper reveals that, in general, condition (1.12) below is sufficient for M to be non-degenerate. ∞ Theorem 1.2. Suppose Assumptions 1-2 hold. If

+ 2 + ∞ 2 + h(y) log h(y)m(dy)+ kpk(y)h(y) log (kh(y))µ(dy) < + , (1.12) E E ∞ Z Z Xk=0 1 then Mt converges to M in L (Px) for every x E, and, consequently, Px(M > 0) > 0. ∞ ∈ ∞ Thus under condition (1.12), X (h) grows exponentially with rate λ . Note that when h is t − 1 bounded, (1.12) is equivalent to (1.11). The next question to ask is that, for a general test function f +(E), what is the limiting behavior of X (f) as t ? By (1.8) and (1.10), it is not hard ∈ B t → ∞ to deduce (see (2.8) below) that for every f +(E) with f ch for some constant c> 0, ∈B ≤ eλ1tE [X (f)] = h(x)P h(f/h)(x) h(x) f, h as t + . x t t → h i → ∞ So, the mean of X (f) also grows exponentially with rate λ . Our previous question is related to t − 1 the question: for f +(E) with f ch for some constant c> 0, does X (f) grow exponentially ∈ B ≤ t with the same rate? If so, can one identify its limit? We first answer these questions in Theorem 1 1.3 and Corollary 1.4 in terms of convergence in L (Px) and in probability, under an additional condition (1.13). h Note that under Assumption 1, for every t> 0, Pt has a symmetric continuous transition density function ph(t,x,y) on E E with respect to the measure m, which is related to p(Q 1)µ(t,x,y) by × − the following formula: p(Q 1)µ(t,x,y)e ph(t,x,y)= eλ1t − , x,y E. h(x)h(y) ∈ Theorem 1.3 (Weak law of large numbers). Suppose Assumptions 1-2 and (1.12) hold. If there exists some t0 > 0 such that

(Q 1)µ h p − (t ,y,y)m(dy) < + , or equivalently, p (t ,y,y)m(dy) < + , (1.13) 0 ∞ 0 ∞ ZE ZE 6 e then for any x E and any f +(E) with f ch for some c> 0, we have ∈ ∈B ≤

λ1t 1 lim e Xt(f)= M f, h in L (Px). t + ∞ → ∞ h i Corollary 1.4. Under the assumptions of Theorem 1.3, it holds that

Xt(f) M lim = ∞ in probability with respect to Px, t + E [X (f)] h(x) → ∞ x t for every x E and every f +(E) with f ch for some c> 0. ∈ ∈B ≤ For almost sure convergence result, we need a stronger condition (1.14) below.

Theorem 1.5 (Strong law of large numbers). Suppose Assumptions 1-2 and (1.12) hold. If there exists t1 > 0 such that p(Q 1)µ(t ,y,y) − 1 or, equivalently, h sup 2 < + , sup p (t1,y,y) < + , (1.14) y E h(y) ∞ y E ∞ ∈ ∈ then there exists Ω0 Ω of P -full probability for every x E, such that, for every ω Ω0 and ⊂ x ∈ ∈ every f (E) with compact support whose set of discontinuous points has zero m-measure, we ∈ Bb have

λ1t lim e Xt(f)(ω)= M (ω) f, h . (1.15) t + ∞ → ∞ h i

Corollary 1.6. Suppose the assumptions of Theorem 1.5 hold and let Ω0 be defined in Theorem 1.5. Then Xt(f)(ω) M (ω) lim = ∞ t + E [X (f)] h(x) → ∞ x t for every ω Ω0 and for every f (E) with compact support whose set of discontinuous points ∈ ∈Bb has zero m-measure.

Remark 1.7. The condition (1.13) is equivalent to

ph(t /2,x,y)2m(dy)m(dx) < + . 0 ∞ ZE ZE Hence by [7, Page 156], P h is a compact operatore on Le2(E, m) for every t t /2. Consequently t ≥ 0 Assumption 2 is automatically satisfied if either (1.13) or (1.14) holds. e To understand condition (1.14), we give some equivalent statements of (1.14) under our As- sumptions 1-2.

Proposition 1.8. Suppose Assumptions 1-2 hold. The following are equivalent to (1.14).

(i) There exists t1 > 0 such that for any t>t1,

h c2t sup p (t,x,y) 1 c1e− (1.16) x,y E | − |≤ ∈

for some c1, c2 > 0.

7 (ii) ph(t,x,y) 1, as t + uniformly in (x,y) E E. (1.17) → ↑ ∞ ∈ ×

(iii) There exist constants t, ct > 0 such that

(Q 1)µ p − (t,x,y) c h(x)h(y) for every x,y E. (1.18) ≤ t ∈

Property (1.18) is called asymptotically intrinsically ultracontractive (AIU) by Kaleta and L˝orinczi in [15]. If the inequality in (1.18) is true for every t > 0, and every x,y E, then (Q 1)µ ∈ P − : t > 0 is called intrinsically ultracontractive (IU). It is shown in [15] that in case of { t } symmetric α-stable processes (α (0, 2)), (AIU) is a weaker property than (IU). ∈

1.4 Examples

In this subsection, we illustrate our main results by several concrete examples. For simplicity, we consider binary branching only, i.e., every particle gives birth to precisely two children, in which + + case Q(x) 2 on E. Since ∞ kp (x) log k 2 log 2 on E, condition (1.12) is reduced to ≡ k=0 k ≡ P h(y)2 log+ h(y) (m(dy)+ µ(dy)) < + . (1.19) ∞ ZE Example 1. [WLLN for branching OU processes with a quadratic branching rate d function] Let E = R . In Example 10 of [11], (X, Πx) is an Ornstein-Ulenbeck (OU) process on Rd with infinitesimal generator 1 = σ2∆ cx on Rd, L 2 − ·∇

c d/2 c x 2 where σ, c> 0. Without loss of generality, we assume σ = 1. Let m(dx)= π e− | | dx. Then X is symmetric with respect to the probability measure m, and the Dirichlet form ( , ) of X on  E F L2(Rd,m) is given by

= f L2(Rd,m) : f(x) 2m(dx) < + , F ∈ d |∇ | ∞  ZR  1 (u, u)= f(x) 2m(dx). E 2 d |∇ | ZR Let β(x) = b x 2 + a with a, b > 0 be the branching rate function. Let P β be the corresponding | | t Feynman-Kac semigroup,

t β Pt f(x) := Πx exp β(Xs)ds f(Xt) .  Z0   Suppose c> √2b and α = √c2 2b. Let − λ := inf λ R : there exists u> 0 such that ( + β λ)u =0 in Rd c { ∈ L − }

8 be the generalized principal eigenvalue. Let φ denote the corresponding ground state, i.e., φ > 0 such that ( +β λ )φ = 0. As indicated in [11], λ = 1 (c α)+a> 0 and φ(x)= α d/4 exp 1 (c L − c c 2 − c { 2 − 2 µ λct β Rd α) x . Note that φ and φ = e− Pt φ on . It is easy to see that in this example  λ1 = λc | | } ∈ F h h − and h(x) = φ(x). The transformed process (X , Πx) is also an OU process with infinitesimal generator 1 = ∆ αx on Rd. L 2 − ·∇ 2 α d/2 α x 2 Note that its invariant probability measure is m(dx) = h(x) m(dx) = π e− | | dx. Let h h p (t,x,y) be the transition density of X with respect to m. It is known that  e 1 d/2 α ph(t,x,y) = exp e x 2 + y 2 2x yeαt 1 e 2αt −(e2αt 1) | | | | − ·  − −   −   In particular, 1 d/2 2α ph(t,x,x)= exp x 2 . 1 e 2αt eαt + 1| |  − −    Thus ph(t,x,x)m(dx) < + for t> 0. Moreover, we observe that condition (1.19) is satisfied Rd ∞ for thisR example. Therefore Theorem 1.3 holds for this example. This example doese not satisfy the assumptions in [6]. To be more specific, here the ground state h is unbounded and β(x)= b x 2 + a is not in the Kato class K (X) of X. | | ∞ Example 2. [WLLN for branching Hunt processes with a bounded branching rate func- tion] Let E be a locally compact separable metric space and m a positive Radon measure on E with full support. Suppose the branching rate function β is a non-negative bounded function on E. Sup- pose the underlying Hunt process (X, Πx) satisfies that for every t> 0, there exists a family of jointly continuous, symmetric and positive kernels p(t,x,y) such that Ptf(x)= E p(t,x,y)f(y)m(dy), and that there exists s1 > 0 so that R p(s ,x,x)m(dx) < + . (1.20) 1 ∞ ZE In this case the Feyman-Kac semigroup t β Pt f(x) := Πx exp β(Xs)ds f(Xt)  Z0   has a jointly continuous and positive kernel pβ(t,x,y). It is easy to see that

β ∞t β β ∞t e−k k p(t,x,y) p (t,x,y) ek k p(t,x,y) for every t> 0 and x,y E. (1.21) ≤ ≤ ∈ Properties (1.20) and (1.21) imply that pβ(s ,x,x)m(dx) < + . Thus P β is a compact operator E 1 ∞ t on L2(E,m) for every t s . By Jentzch’s theorem (see, for example, [22, Theorem V.6.6]), λ ≥ 1 R − 1 is a simple eigenvalue of + β where is the infinitesimal operator of X, and an eigenfunction h L L of + β associated with λ can be chosen to be positive and continuous on E. Suppose λ < 0. L − 1 1 We assume in addition that there exists s2 > 0 such that

p(s ,x,x)2m(dx) < + . (1.22) 2 ∞ ZE 9 β It follows from (1.21) and H¨older’s inequality that for every t>s2, Pt is a bounded operator from L2(E,m) to L4(E,m). Thus h = eλ1tP βh L4(E,m) and so condition (1.19) is satisfied. Hence t ∈ Theorem 1.3 holds. Conditions (1.20) and (1.22) are satisfied by a large class of Hunt processes, which contains subordinated OU processes as special cases. By “subordinated OU process” we mean the process d Xt = YSt , where Yt is an OU process on R and St is a subordinator on R+ independent of Yt. In the special case S t, X reduces to the OU process. In general, the sample path of X is t ≡ t t discontinuous. Suppose the infinitesimal generator of Yt is 1 = σ2∆ bx on Rd L 2 − ·∇ where σ,b > 0 are constants, andb St is a subordinator with positive drift coefficient a > 0.

As is indicated in Example 1, Yt is symmetric with respect to the reference measure m(dx) := d/2 b exp b x 2/σ2 dx. We usep ˆ(t,x,y) to denote the transition density of Y with respect to πσ2 {− | | } t m. It is known that d/2 2bt − b 2 2 bt pˆ(t,x,y) = 1 e− exp x + y 2x ye . − −σ2 (e2bt 1) | | | | − ·    −   By definition, the transition density of Xt with respect to m is given by

p(t,x,y)= E [ˆp(St,x,y)] .

It is proved in [20, Example 1.1] that (1.20) and (1.22) hold for a subordinated OU process. Therefore, Theorem 1.3 holds for branching subordinated OU processes with a bounded branching rate function.

Example 3. [LLN for branching diffusions on bounded domains with branching rate given by a Kato class measure] Suppose d 3, E Rd is a bounded C1,1 domain (that is, the ≥ ⊂ boundary of E can be locally characterized by C1,1 functions) and m is the Lebesgue measure on E. Let d 1 = ∂ (a ∂ ) L 2 i ij j i,jX=1 with a (x) C1(Rd) for every i, j = 1, , d. Suppose the matrix (a (x)) is symmetric and ij ∈ · · · ij uniformly elliptic. It is known that there exists a symmetric diffusion process Y on Rd with generator . Let X be the killed process of Y upon E, i.e., L

Yt, t<τE, Xt = ( ∂, t τ , ≥ E where τ := inf t > 0 : Y E and ∂ is a cemetery state. Then X has a transition density E { t 6∈ } function pE(t,x,y) which is jointly continuous in (x,y) and positive for every t> 0. The following two-sided estimates of pE(t,x,y) is established in [21, Theorem 2.1], extending an earlier result of Q. Zhang. Let f (t,x,y) := 1 δE (x) 1 δE (y) , where δ (x) denotes the distance between E ∧ √t ∧ √t E     10 x and the boundary of E. There exist positive constants c , i = 1, , 4, such that for every i · · · (t,x,y) (0, 1] E E, ∈ × × 2 2 d/2 c2 x y /t d/2 c4 x y /t c f (t,x,y)t− e− | − | p (t,x,y) c f (t,x,y)t− e− | − | . 1 E ≤ E ≤ 3 E We say that a signed smooth Radon measure ν on Rd belongs to the Kato class K (α (0, 2]) if d,α ∈ ν (dy) lim sup | | d α = 0. (1.23) r 0 x Rd x y r x y − ↓ ∈ Z| − |≤ | − | d In fact Kd,α is the Kato class of the rotationally symmetric α-stable processes on R . We assume the branching rate measure µ is a non-negative Radon measure in K . For any f +(E), let d,2 ∈B µ µ Pt f(x):=Ex [exp(At )f(Xt)] .

µ µ Then Pt has a transition density pE(t,x,y) which is jointly continuous in (x,y) and positive for every t > 0. It is shown in [16, Theorem 4.4] that there exist positive constants c , i = 5, , 8, i · · · such that for every (t,x,y) (0, 1] E E, ∈ × × 2 2 d/2 c6 x y /t µ d/2 c8 x y /t c f (t,x,y)t− e− | − | p (t,x,y) c f (t,x,y)t− e− | − | . (1.24) 5 E ≤ E ≤ 7 E The infinitesimal generator of P µ is ( + µ) with zero Dirichlet boundary condition. It follows t L |E from Jentzch’s theorem that λ is a simple eigenvalue of ( + µ) and that an eigenfunction h − 1 L |E associated with λ can be chosen to be positive with h 2 = 1. Immediately, h is continuous − 1 k kL (E,dx) on E by the dominated convergence theorem. We assume λ1 < 0. Recall that E is bounded. Using µ the equation h = eλ1 P h and the estimates in (1.24), we get that for every x E, 1 ∈ c (1 δ (x)) h(x) c (1 δ (x)) (1.25) 9 ∧ E ≤ ≤ 10 ∧ E for some positive constants c9, c10. Let

eλ1tpµ (t,x,y) ph (t,x,y) := E for x,y E. E h(x)h(y) ∈

Immediately condition (1.19) holds by the boundedness of h and condition (1.14) holds by (1.24) and (1.25). Therefore both Theorem 1.3 and Theorem 1.5 hold for this example.

Example 4. [LLN for branching killed α-stable processes with a bounded branching rate function] Suppose d 1, E = Rd, m is the Lebesgue measure on Rd and α (0, 2). Suppose ≥ d ∈ Y is a symmetric α- on R , and c(x) is a non-negative function in Kd,α ( a function q is said to be in Kd,α, if the measure ν(dx) := q(x)dx is in Kd,α where Kd,α is defined in (1.23)). Let X be the subprocess of Y such that for all f (Rd), ∈Bb t P f(x):=E (f(X ))=E exp c(Y )ds f(Y ) . t x t x − s t   Z0  

11 It is known that the infinitesimal generator of X is = ∆α/2 c(x), where ∆α/2 := ( ∆)α/2 is L − − − the generator of a symmetric α-stable process. Let the branching rate function β be a non-negative bounded function on Rd. Let V (x) := c(x) β(x). Clearly, V K . For any f +(Rd), let − | |∈ d,α ∈B t β Pt f(x):=Ex exp β(Xs)ds f(Xt) .  Z0   1 d d Note that for every t > 0, Pt is bounded from L (R , dx) to L∞(R , dx), and satisfies the strong β p1 d p2 d Feller property. It follows from [1] that for very t> 0, Pt is bounded from L (R , dx) to L (R , dx) for any 1 p p + . Thus under our Assumption 1, the ground state h is a positive bounded ≤ 1 ≤ 2 ≤ ∞ continuous function on Rd. The semigroup P β : t 0 is the Feynman-Kac semigroup with { t ≥ } infinitesimal generator β = ∆α/2 V . Assume in addition that V (x) = c(x) β(x) satisfies the L − + − following conditions: lim inf x + V (x)/ log x > 0 and V − has compact support. Then by [15, | |→β ∞ | | Theorem 5.5], the semigroup Pt is (AIU). Hence both Theorem 1.3 and Theorem 1.5 are true for this example. It is known from [15] that in case of symmetric α-stable processes, (AIU) is weaker than (IU). For instance, β(x)= c(x) β(x) with c(x) = log x 1 x R and β has compact support − | | {| |≥ } in B(0,R) for some R 1 is such an example. ≥ The rest of the paper is organized as follows. In Section 2, we review some facts on Girsanov transform and h-transforms in the context of symmetric Markov processes, and prove Proposition 1.8. Spine construction of branching processes is recalled in Section 3. We then present proof for the L log L criteria, Theorem 1.2, in Section 4. Weak law of large numbers, Theorem 1.3, is proved in Section 5, while Theorem 1.5 on the strong law of large numbers will be proved in Section 6. The lower case constants c , c , , will denote the generic constants used in this article, whose 1 2 · · · exact values are not important, and can change from one appearance to another.

2 Preliminary

Recall h µ is the minimizer in Assumption 1. Since h , by Fukushima’s decomposition, we ∈ F ∈ F have for q.e. x E, Π -a.s. ∈ x h(X ) h(X )= M h + N h, t 0, t − 0 t t ≥ h h where M is a martingale additive functional of X having finite energy and Nt is a continuous h additive functional of X having zero energy. It follows from (1.7) and [13, Theorem 5.4.2] that Nt is of bounded variation, and

t t h (Q 1)µ N = λ h(X )ds h(X )dA − , t 0. t − 1 s − s s ∀ ≥ Z0 Z0 Following [4, Section 2] (see also [6, Section 2]), we define a on the random time interval [0,ζp) by t 1 h Mt := dMs , t [0,ζp), (2.1) 0 h(Xs ) ∈ Z −

12 where ζp is the predictable part of the life time ζ of X. Then the solution Rt of the stochastic differential equation t Rt =1+ Rs dMs, t [0,ζp), (2.2) − ∈ Z0 is a positive local martingale on [0,ζp) and hence a supermartingale. As a result, the formula

dΠh = R dΠ on t<ζ for x E x t x Ht ∩ { } ∈ uniquely determines a family of subprobability measures Πh : x E on (Ω, ). We denote X { x ∈ } H under Πh : x E by Xh, that is { x ∈ } h h Πx f(Xt ) = Πx [Rtf(Xt) : t<ζ] h i for any t 0 and f +(E). It follows from [4, Theorem 2.6] that the process Xh is an irreducible ≥ ∈B recurrent m-symmetric right Markov process, where m(dy) := h(y)2m(dy). Note that by (2.1), (2.2) and Dol´ean-Dade’s formula, e e 1 c h(Xs) h(Xs) Rt = exp Mt M t exp 1 , t [0,ζp), (2.3) − 2h i h(Xs ) − h(Xs ) ∈ 0

1 c h(Xs) h(Xs) h(Xs ) (Q 1)µ log h(Xt) log h(X0)= Mt M t + log − − λ1t At − . (2.4) − − 2h i h(Xs ) − h(Xs ) − − s t  − −  X≤ By (2.3) and (2.4), we get

(Q 1)µ h(Xt) Rt = exp λ1t + At − on [0,ζ). h(X0)  Therefore for any f +(E), ∈B

λ1t λ1t h h e A(Q−1)µ e (Q 1)µ h Π f(X ) = Π e t h(X )f(X ) = P − (hf)(x)= P f(x). (2.5) x t h(x) x t t h(x) t t h i h i Let ( h, h) be the symmetric Dirichlet form on L2(E, m) generated by Xh. Then (2.5) says that E F (Q 1)µ the transition semigroup of Xh is exactly the semigroup P h : t 0 obtained from P − { t ≥ } t through Doob’s h-transform. Consequently, f h if ande only if fh µ, and ∈ F ∈ F h (Q 1)µ 2 2 (f,f)= − (fh, fh) λ f(x) h(x) m(dx). E E − 1 ZE In other words, Φh : f fh is an isometry from ( h, h) onto ( (Q 1)µ+λ1m, µ) and from 7→ E F E − F L2(E, m) onto L2(E,m). Let σ( h) denote the spectrum of the positive definite self-adjoint operator E associated with h. We know from [4, Theorem 2.6] that the constant function 1 belongs to h, E F e

13 and h(1, 1) = 0. Hence 0 σ( h) is a simple eigenvalue and 1 is the corresponding eigenfunction. E ∈ E Therefore λh := inf h(u, u) : u h with u(x)2m(dx) = 1 = 0. 1 E ∈ F  ZE  Let λh be the second bottom of σ( h), i.e. 2 E e

λh := inf h(u, u) : u h with u(x)m(dx)=0 and u(x)2m(dx) = 1 . 2 E ∈ F  ZE ZE  h h e e In view of the isometry Φ , we have λ2 = λ2 λ1. So Assumption 2 is equivalent to assuming h − λ2 > 0. Define a (x) := ph(t,x,x) for t> 0 and x E. (2.6) t ∈ Note that by (1.9) for any x E and t,s 0, e∈ ≥ P h ϕ(x) = P hP hϕ(x) | t+s | | s t | ph(s,x,y) P hϕ(y) m(dy) ≤ | t | ZE 1/2 h 2 h p (s,x,y) m(dy)e P ϕ 2 e ≤ k t kL (E,m) ZE  1/2 λht a (x) e− ϕ 2 e . (2.7) ≤ 2s k ekL (E,m) 2 We use f, g 2 e to denote f(x)g(x)m(dx). For every g L (E, m) we can decompose g as h iL (m) E e ∈ g(x)= g, 1 2 e + ϕ(x) with ϕ(x) satisfying ϕ(x)m(dx) = 0. Then by (2.7), for any t>s 0 h iL (m) R E ≥ and any x E, e e ∈ R e h h λh(t s) 1/2 P g(x) g, 1 2 e = P ϕ(x) e− − a (x) ϕ 2 e | t − h iL (m)| | t |≤ 2s k kL (E,m) λh(t s) 1/2 e− − a2s(x) g L2(E,me ). (2.8) ≤ k ke

Proof of Proposition 1.8. We only need to prove thate (1.14) implies (1.16). For any f L2(E, m), ∈ define c := f(x)m(dx). Immediately, we have for any t > 0, (f c )m(dx) = 0, and f E E − f f c L2(E, m). By (1.9), we have e − f ∈ R R e e h λht P f c 2 e e− f 2 e . (2.9) e k t − f kL (E,m) ≤ k kL (E,m) Let t > 0 be the constant in (1.14). By the semigroup property, for any t t /2 and x E, 1 ≥ 1 ∈ ph(t,x,y) = ph(t /2,x,z)ph(t t /2,y,z)m(dz) 1 − 1 ZE = P h f (y), (2.10) t t1/2 x − e where f ( ) := ph(t /2, x, ). Note that by H¨older’s inequality, x · 1 · ph(t,x,y) 1 = ph(t/2,x,z)ph(t/2,z,y)m(dz) 1 | − | | − | ZE 14 e = (ph(t/2,x,z) 1)(ph(t/2,z,y) 1)m(dz) | − − | ZE 1/2 1/2 (ph(t/2,x,z) 1)2m(dz) (eph(t/2,z,y) 1)2m(dz) (2.11). ≤ − − ZE  ZE  h 2 Note that cfx = E p (t1/2,x,y)m(dy) = 1 and eE fx (y)m(dy) = at1 (x). By (2.10)e and (2.9), for any t>t1, R R e 1/2 e e h 2 h (p (t/2,x,z) 1) m(dz) = P fx c 2 e − k (t t1)/2 − fx kL (E,m) ZE  − λh(t t1)/2 e− − f 2 e e ≤ k xkL (E,m) λh(t t1)/2 1/2 = e− − at1 (x) . (2.12) Similarly, for any t>t , 1 e 1/2 h 2 λ (t t1)/2 1/2 (p (t/2,z,y) 1) m(dz) e− h − a (y) . (2.13) − ≤ t1 ZE  Combining (2.11), (2.12) and (2.13), we havee for any t>t1, e

h λ (t t1) 1/2 1/2 p (t,x,y) 1 e− h − a (x) a (y) . (2.14) | − |≤ t1 t1 Therefore (1.14) implies that for any t>t , (1.16) holds. 1 e e

3 Spine construction

1 To establish the L convergence of the martingale Mt, we apply the “spine” and change of measure techniques presented in [14] for branching diffusions to our branching Hunt processes. The notation used here is closely related to that used in [14]. It is known that the family structure of the particles in a branching Markov process can be well expressed by marked Galton-Watson trees (see, for example, [14, 19] and the references therein). Let denote the space of all marked G-W trees. For T a fixed τ , all particles in τ are labeled according to the Ulam-Harris convention, for example, ∈T 231 or 231 is the first child of the third child of the second child of the initial ancestor . Besides, ∅ ∅ each particle u τ has a mark (X ,σ , A ) where X : [b ,ζ ) E is the spatial location of u ∈ u u u u u u → ∂ during its life span (b is its birth time and ζ its fission time), σ = ζ b is the length of its life u u u u − u span, and A denotes the number of its offspring. We use u v to mean that u is an ancestor of v. u ≺ Since in this paper every particle is assumed to give birth to at least one child, for each tree τ, we can choose a distinguished genealogical path of descent from the initial ancestor . Such a line ∅ is called a spine and denoted as ξ = ξ = ,ξ ,ξ , , where ξ is the label of the ith spine node. { 0 ∅ 1 2 ···} i Define node (ξ) := u if u ξ and is alive at time t. We shall use X : t 0 and n : t 0 t ∈ { t ≥ } { t ≥ } respectively to denote the spatial path and the counting process of fission times along the spine. e Let denote the space of G-W trees with a distinguished spine. We introduce some fundamental T filtrations that encapsulate different knowledge: e t := σ (u, Xu,σu, Au),ζu t; (u, Xu(s),s [bu,t]),t [bu,ζu) , := t := σ t; t 0 ; F { ≤ ∈ ∈ } F∞ F {F ≥ } t 0 _≥ 15 t := σ t, nodet(ξ) , = t; F {F } F∞ F t 0 _≥ e e e t := σ Xs : s t , := t; G { ≤ } G∞ G t 0 _≥ e t := σ t, (nodes(ξ) : s t), (Au : u nodet(ξ)) , := t. G {G ≤ ≺ } G∞ G t 0 _≥ e e e We assume Px is the measure on ( , ) such that the filtered probability space , , ( t)t 0 , Px T F∞ T F∞ F ≥ is the canonical model for the branching Hunt process X described in the introduction. We know e e from [14] that the measure Px on ( , ) can be extended to the probability measure Px on T F∞ ( , ) such that the nth spine node is uniformly chosen from the children of the (n 1)th spine T F∞ e − e node. e e For every t 0, as in [14], we define ≥

(Q 1)µ h(Xt) λ t Xt(h) λ t h(Xt) η(t) := exp λ t + A − ,Z(t) := e 1 , and η(t) := e 1 A . 1 t h(x) h(x) h(x) v v nodet(ξ) n o e e ≺ Y e Then η(t), Z(t) and η(t) are positive P -martingales with respect to the filtrations , and , x Gt Ft Ft respectively. Moreover, both η(t) and Z(t) are projections of η(t) onto their filtrations: e e e Z(t)= P (η(t) ) , η(t)= P (η(t) ) for t 0. x |Ft x |Gte ≥ We call η(t) the single-particlee martingalee and Z(t)e theebranching-particle martingale. As in [14], we define a new probability measure Qx by setting

dQ ee = η(t)dP e for t 0. (3.1) x t x t |F |F ≥ This implies that e e e

dQx t = Z(t)dPx t for t 0. (3.2) |F |F ≥ The influence of the measure changee (3.1) liese in the following three aspects: Firstly, under Qx the motion of the spine is biased by the martingale η(t). Secondly, the branching events along e the spine occur at an accelerated rate. Finally, the number of children of the spine nodes is size- biased distributed, that is, for every spine node v located at x, Av is distributed according to the probability mass function p (x) := kp (x)/Q(x) : k = 0, 1, , while other (non-spine) nodes, { k k ···} once born, remain unaffected. More specifically, under measure Qx, b h h (i) The spine’s spatial process X moves in E as a copy of (X e, Πx);

(ii) The branching events alonge the spine occur at an accelerated rate µ(dx) := Q(x)h(x)2µ(dx). This implies that given , for every t > 0, the number of fission times nt is a Poisson G∞ random variable with parameter At, where At is a PCAF of X havinge Revuz measure µ. To µe emphasize this dependence, we also write At for At. e e e e e 16 (iii) At the fission time of node v in the spine, the single spine particle is replaced by Av children,

with Av being distributed according to the size-biased distribution pk(Xζv ) : k = 0, 1, . { − ···} Each child is selected to be the next spine node with equal . b e (iv) Each of the remaining A 1 children gives rise to independent subtrees, which are not part v − of the spine and evolve as independent processes determined by the measure P shifted to their point and time of creation.

For more details on martingale changes of measures for branching Hunt processes, see [14] and [19].

4 L log L criteria

In this section, we prove Theorem 1.2. It follows from [5, Theorem 4.1.1] that if ν is a smooth measure of Xh, then for every g +(E) and t 0, ∈B ≥ t t h h ν Πme g(Xs )dAs = g(y)ν(dy)ds. (4.1) Z0  Z0 ZE Since t is arbitrary, the monotone class theorem implies that for any f(s,x) = l(s)g(x) with l +[0, + ) and g +(E), ∈B ∞ ∈B t t h h ν Πme f(s, Xs )dAs = f(s,y)ν(dy)ds. (4.2) Z0  Z0 ZE Note that for every f +([0, + ) E), there exists a sequence of functions f (s,x)= l (s)g (x) ∈B ∞ × n n n with l +[0, + ) and g +(E) such that f (s,x) converges increasingly to f(s,x) for all s n ∈B ∞ n ∈B n and x. Thus by the monotone convergence theorem, (4.2) holds for all f +([0, + ) E). ∈B ∞ ×

Proof of Theorem 1.2. Since Mt is a positive martingale, it suffices to prove that ExM = h(x) for ∞ all x E. By [8, Theorem 5.3.3] (or [19, Lemma 3.1]), it suffices to show that if (1.12) holds, then ∈

Qx lim sup Mnσ < + = 1 for every σ > 0 and x E. (4.3) N n + ∞ ∈  ∋ → ∞  e Recall that contains all the information about the spine. We have the following spine decom- G∞ position for Mt, e

λ1t Qx Mt = Qx e h(Xu(t)) G∞ " G∞# h i u t X∈Z e e eλ1t e λ1ζu λ1(t ζu) = e h(Xt)+ (Au 1)e E e e − Xt ζu (h) − Xζu − u ξn X≺ t h i e = eλ1th(X )+ (A 1)eλ1ζu h(X ). (4.4) t u − ζu u ξn X≺ t e e

17 Let := σ X : s t and be the trivial σ-field. It follows from the second Borel-Cantelli Gt { s ≤ } G0 lemma (see, for example, [8, Section 5.4]) and the that for any σ> 0 and M 1, e ≥ + ∞ λ1nσ λ1nσ Qme e h(Xnσ) > M i.o. = Qme Qme e h(Xnσ) > M (n 1)σ =+ G − ∞!   nX=1   + e e e ∞ e e h h λ1nσ = Πme Π e e h(Xσ) > M =+ . (4.5) X(n−1)σ ∞! nX=1   h e Recall that m is the invariant probability measure of (X, Πx). Thus by Fubini’s theorem and Markov property, we have e e + + ∞ ∞ h h λ1nσ h λ1nσ Πme Π e e h(Xσ) > M = Πme e h(Xnσ) > M X(n−1)σ "n=1 # n=1 X   X   + e ∞ e = 1 eλ1nσh(y)>M m(dy) E { } nX=1 Z + ∞ log+ h(y) logeM = m − > nσ λ1 nX=1  −  1 + ( λe)− log h(y)m(dy) < . ≤ − 1 ∞ ZE λ1nσ Therefore by (4.5) we have Qme e h(Xnσ) > M i.o. = 0. Consequently, e   e e λ1nσ Qm lim sup e h(Xnσ) < + = 1. N n + ∞  ∋ → ∞ 

e e λ1nσ It is easy to check that the function x Qx lim supN n + e h(Xnσ) < + is an invariant 7→ ∋ → ∞ ∞ h function for (X, Π e ). Recall that X has a transition density function with respect to m. By [5, m e e Theorem A.2.17], e e e λ1nσ Qx lim sup e h(Xnσ) < + = 1 for every x E. N n + ∞ ∈  ∋ → ∞  Suppose ε (0, λ ).e For simplicity we usee ζ and A to denote respectively the fission time and ∈ − 1 i i offspring number of the ith spine node. + + + ∞ ∞ ∞ λ1ζi λ1ζi λ1ζi e A h(X ) = e A h(X )1 e εζ + e A h(X )1 e εζ i ζi i ζi Aih(X ) e i i ζi Aih(X )>e i { ζi ≤ } { ζi } Xi=1 Xi=1 Xi=1 e =: I + II. e e (4.6)

Recall that contains all the information of spine’s motion Xt : t 0 in E. For any set G∞ { ≥ } B [0, + ) (Z+), define N(B) := ♯ i 1 : (ζi, Ai) B . Then conditioned on , ∈ B ∞ × B { ≥ ∈ e }µe G∞ N is a Poisson random measure on [0, + ) Z with intensity dA pˆ (X )δ (dy), where + s k Z+ k s k ∞ × ∈ µ(dx)= Q(x)h(x)2µ(dx). We have P e + + + ∞ h ∞ ∞ µe e Qx 1 e εζ = Π pˆk(ξs)1 e εs dA . Aih(Xζ )>e i x kh(Xs)>e s { i }! 0 { } ! Xi=1 Z Xk=0 e 18 Note that by (4.2) and our assumption (1.12),

+ + + ∞ h ∞ ∞ µe Q e 1 e εζ = Π e pˆ (ξs)1 e εs dA m Aih(Xζ )>e i m k kh(Xs)>e s { i }! 0 { } ! Xi=1 Z Xk=0 e + + ∞ ∞ = pˆk(y)1 log+(kh(y))>εs µ(dy)ds 0 E { } Z Z Xk=0 + 1 ∞ + e2 = ε− kpk(y) log (kh(y))h(y) µ(dy) < + . E ∞ Z Xk=0 Thus + ∞ Q 1 e εζ < + =1 for m-a.e. x E. x Aih(X )>e i { ζi } ∞! ∈ Xi=1 e This implies that II is the sum of a finite many terms and so Qe(II < + )=1 for m-a.e. x E. x ∞ ∈ On the other hand, since Q(x) is bounded on E, we have e e + + ∞ h ∞ λ1s µe εs Qme (I) = Πme e pˆk(Xs)kh(Xs)1 kh(ξs) e dAs " 0 { ≤ } # Z Xk=0 + e + ∞ e e ∞ λ1s = e pˆk(y)kh(y)1 kh(y) eεs µ(dy)ds 0 E { ≤ } Z Z Xk=0 + + ∞ e ∞ (λ1+ε)s 2 e pˆk(y)Q(y)h(y) µ(dy)ds ≤ 0 E Z Z Xk=0 Q h(y)2µ(dy) < + . ≤ k k∞ ∞ ZE Thus we have Q (I < + )=1 for m-a.e. x E. Now we have proved that x ∞ ∈ + ∞ e λ1ζi e Qx e Aih(Xζi ) < + =1 for m-a.e. x E. " ∞# ∈ Xi=1 e e e + λ ζ It is easy to check that the function x Q ∞ e 1 i A h(X ) < + is an invariant function. 7→ x i=1 i ζi ∞ + λ ζ Thus by [5, Theorem A.2.17], we get Q hP∞ e 1 i A h(X ) < + =i 1 for every x E. By xe i=1 i ζie ∞ ∈ (4.4) we have h i e P e Qx lim sup Qx Mnσ < + = 1 for every x E. N n + |G∞ ∞ ∈  ∋ → ∞    e e e 1 By Fatou’s lemma, we get Qx (lim infN n + Mnσ < + ) = 1. Note that Mnσ− is a positive ∋ → ∞ ∞ Qx-martingale with respect to nσ, so it converges almost surely as n . It follows then e F →∞ e Qx lim sup Mnσ = lim inf Mnσ < + = 1. N n + N n + ∞  ∋ → ∞ ∋ → ∞  e

19 5 Weak law of large numbers

In this section, we present a proof for Theorem 1.3.

Lemma 5.1. If Assumption 1 and (1.12) hold, then for every t 0 and φ +(E), ≥ ∈Bb E X (φh) log+ X (φh) h(x)m(dx) < + . (5.1) x t t ∞ ZE   Proof. First we note that for every x E and φ +(E), ∈ ∈Bb

+ λ1t λ1t 1 + Ex Xt(φh) log Xt(φh) e− φ h(x)Ex e h(x)− Xt(h) log Xt(φh) ≤ k k∞ λ1t h + i   = e− φ h(x)Qx log Xt(φh) . (5.2) k k∞   Recall that contains all the information about the spine.e Since for any a, b 0 G∞ ≥ e log+(a + b) log+ a + log+ b + log 2, (5.3) ≤ we have by Jensen’s inequality

+ + Qx log Xt(φh) = Qx Qx log Xt(φh) |G∞   h  i e Qe x loge Qx Xt(φh) e1 ≤ ∨ |G∞ h +  i Qe x log e Qx Xt(φh) e + 1 ≤ |G∞ h +    i Qe x log Qex Xt(φh) e + log 2. (5.4) ≤ |G∞ h  i By (5.2) and (5.4), it suffices to show that e e e

+ 2 Qx log Qx Xt(φh) h(x) m(dx) < + . (5.5) E |G∞ ∞ Z h  i e e e We get the spine decomposition for Qx Xt(φh) as follows: |G∞   e e Qx Xt(φh) = (φh)(Xt)+ (Au 1)E e [Xt ζu (φh)] |G∞ − Xζu − u ξt   X≺ e e e λ1(t ζu) (φh)(Xt)+ Au φ e− − h(Xζu ). (5.6) ≤ k k∞ u ξt X≺ e e Note that log+(ab) log+ a + log+ b for any a, b > 0. Using this and an analogy of (5.3) as well as ≤ the assumption that λ1 < 0, we have

+ log Qx Xt(φh) |G∞   + e e λ1(t ζu) log φh(Xt)+ Au φ e− − h(Xζu ) ≤  k k∞  u ξt X≺ +  e + λ1(t eζu)  + log (φh(Xt)) + log Au φ e− − h(Xζu ) + log nt ≤ k k∞ u ξt X≺   e e 20 + + + log (φh(Xt)) + log φ λ1(t ζu) + log Auh(Xζu ) + nt ≤ k k∞ − − u ξt X≺    + e + + e log (φh(Xt)) + log Auh(Xζu ) + (1 + log φ λ1t)nt. (5.7) ≤ k k∞ − u ξt X≺   e e By (1.12) and using the fact that m(dy) is an invariant distribution,

+ 2 + Qx log (φh(Xte)) h(x) m(dx)= log (φh)(x)m(dx) < + . E E ∞ Z   Z Hence (5.5) is impliede by e e

+ 2 Qx log (Auh(Xζu )) + nt h(x) m(dx) < + . (5.8) E   ∞ Z u ξt X≺ e  e  Recall that conditioned on , N( ) = ♯ i 1 : (ζi, Ai) is a Poisson random measure on G∞ e · { ≥ ∈ ·} [0, + ) Z with intensity dAµ p (X )δ (dy). We have + s k Z+ k s k ∞ × ∈ P e + b 2 Qx log (Auh(Xζu )) + nt h(x) m(dx) E   Z u ξt X≺ e  e  + 2 = Qx Qx log (Auh(ξζu )) + nt h(x) m(dx) E   G∞ Z u ξt X≺ e e t +  h ∞ + µe = Πx pk(Xs) log (kh(Xs)) + 1 dAs m(dx) E " 0 # Z Z Xk=0   t + e e ∞ b + 2 e = ds pk(y) log (kh(y)) + 1 Q(y)h(y) µ(dy) 0 E Z Z Xk=0 +  ∞ b + 2 2 t kpk(y) log (kh(y))h(y) µ(dy)+ Q h(y) µ(dy) . ≤ E k k∞ E ! Z Xk=0 Z Immediately the last term is finite by (1.12). Hence we complete the proof.

Lemma 5.2. If Assumptions 1-2 and (1.12) hold, then for any s,σ > 0, m N, and any x E, ∈ ∈

λ1(s+t) λ1(s+t) 1 lim e Xs+t(φh) Ex e Xs+t(φh) t = 0 in L (Px), (5.9) t + → ∞ − F h i λ1(n+m)σ λ1(n+m)σ lim e X(n+m)σ(φh) Ex e X(n+m)σ(φh) nσ = 0 Px-a.s. (5.10) N n + − F ∋ → ∞ + h i for every φ (E). ∈Bb Proof. For any particle u , let Xu,s : t s denote the branching Markov process initiated by ∈Zs { t ≥ } u at time s. It is known that conditioned on , Xu,s and Xv,s are independent for every u, v Fs ∈Zs with u = v. For every s,t 0, define 6 ≥ λ1t λ1t u,t Ss,t := e Xs+t(φh)= e Xs+t(φh), u t X∈Z 21 and λ1t u,t X u,t Ss,t := e s+t(φh)1 X (φh) e−λ1t . { s+t ≤ } u t X∈Z e Obviously S S . s,t ≥ s,t First by the conditional independence and the Markov property we have e 2 E S E (S ) x mσ,nσ − x mσ,nσ Fnσ     = E Vare S e x mσ,nσ Fnσ h  i 2λ1nσ u,nσ E e X u,nσ −λ nσ = e x Var (n+m)σ(φh)1 X (φh) e 1 nσ " { (n+m)σ ≤ } F # u nσ   ∈ZX 2 2λ1nσ u,nσ E E X u,nσ −λ nσ e x (n+m)σ(φh) 1 X (φh) e 1 nσ ≤ { (n+m)σ ≤ } F " u nσ  !# ∈ZX  

2λ1nσ 2 E E X −λ nσ = e x Xu(nσ) ( mσ(φh)) 1 Xmσ (φh) e 1 . (5.11) { ≤ } "u nσ # ∈ZX   2 E X −λ nσ Let gmσ,nσ(y) := y ( mσ(φh)) 1 Xmσ (φh) e 1 . Immediately we have { ≤ }   λ1nσ λ1(n+m)σ gmσ,nσ(y) e− Ey[Xmσ(φh)] e− φ h(y). ≤ ≤ k k∞ Thus g L2(E,m), and mσ,nσ ∈

λ1(n+m)σ gmσ,nσ L2(E,m) e− φ . (5.12) k k ≤ k k∞ Using (2.8), we continue the estimates in (5.11) to get that for n N with nσ > 1, ∈ 2 E S E S x mσ,nσ − x mσ,nσ Fnσ  h i  e2λ1nσE [X (g )] ≤ e x nσ mσ,nσe λ1nσ h = e h(x)Pnσ(gmσ,nσ/h)(x)

λ1nσ λh/2 1/2 λhnσ e h(x) g , h + e a (x) e− g 2 ≤ h mσ,nσ i 1 k mσ,nσkL (E,m) λ1nσ h λh/2 1/2 λhnσ λ1mσ i e h(x) gmσ,nσ , h + e h(x)a1(x) e− − φ . (5.13) ≤ h i e k k∞ Note that e + + ∞ ∞ λ1nσ λ1nσ 2 E X −λ nσ e gmσ,nσ, h = e y ( mσ(φh)) 1 Xmσ (φh) e 1 h(y)m(dy) h i E { ≤ } n=1 n=1 Z h i X X + ∞ λ1(s 1)σ 2 E X −λ sσ h(y)m(dy) e − y ( mσ(φh)) 1 Xmσ (φh) e 1 ds ≤ E 1 { ≤ } Z Z + h i ∞ 1 2 = h(y)m(dy) E (X (φh)) 1 −λ σ dx 2 y mσ Xmσ (φh) xe 1 E 1 λ1σx { ≤ } Z Z − h i + + 1 ∞ 2P X ∞ = h(y)m(dy) z y ( mσ(φh) dz) 2 dx E 0 ∈ 1 zeλ1σ λ1σx Z Z Z ∨ − 22 + 1 λ1σ ∞ e− h(y)m(dy) zP (X (φh) dz) ≤ λ σ y mσ ∈ − 1 ZE Z0 1 λ1σ = e− h(y)E [X (φh)] m(dy) λ σ y mσ − 1 ZE 1 λ1σ λ1mσ 2 e− − φ h(y) m(dy) < + , (5.14) ≤ λ σ k k∞ ∞ − 1 ZE λ1(s 1)σ where in the second equality above we used the change of variables x = e− − . It follows from 2 + (5.13) and (5.14) that ∞ E S E [S ] < + . Thus by the Borel-Cantelli n=1 x mσ,nσ − x mσ,nσ|Fnσ ∞   lemma, P   e e lim Smσ,nσ Ex Smσ,nσ nσ = 0 Px-a.s. (5.15) n + → ∞ − |F Note that for every n,m N, we have h i ∈ e e E S S = E E (S ) E (S ) x mσ,nσ − mσ,nσ x x mσ,nσ|Fnσ − x mσ,nσ|Fnσ h i h i λ1nσ e E E Xe −λ nσ = e x Xu(nσ) mσ(φh)1 Xmσ (φh)>e 1 . (5.16) { } "u nσ # ∈ZX h i

E X −λ nσ Let fmσ,nσ(y) := y mσ(φh)1 Xmσ (φh)>e 1 . Obviously, { } h i λ1mσ fmσ,nσ(y) φ Ey(Xmσ(h)) e− φ h(y) (5.17) ≤ k k∞ ≤ k k∞ and so f L2(E,m). Then by (2.8), we have for n N with nσ > 1, mσ,nσ ∈ ∈ f RHS of (5.16) = eλ1nσE (X (f )) = h(x)P h mσ,nσ (x) x nσ mσ,nσ nσ h   λh/2 1/2 λhnσ h(x) f , h + e a (x) h(x)e− f 2 ≤ h mσ,nσ i 1 k mσ,nσkL (E,m) λh/2 1/2 λhnσ λ1mσ h(x) fmσ,nσ , h + e a1(x) h(x)e− − φ , (5.18) ≤ h i e k k∞ where in the last inequality we used (5.17). Note thate + + ∞ ∞ E X −λ nσ fmσ,nσ, h = y mσ(φh)1 Xmσ (φh)>e 1 h(y)m(dy) h i E { } n=1 n=1 Z   X X+ ∞ E X −λ σs ds y mσ(φh)1 Xmσ (φh)>e 1 h(y)m(dy) ≤ 0 E { } Z Z  + +  ∞ ∞ = h(y)m(dy) ds zPy (Xmσ(φh) dz) −λ σs ∈ ZE Z0 Ze 1 + 1 + = h(y)m(dy) ∞ dt ∞ zP (X (φh) dz) λ σt y mσ ∈ ZE Z1 − 1 Zt 1 + z 1 = h(y)m(dy) ∞ zP (X (φh) dz) dt λ σ y mσ ∈ t − 1 ZE Z1 Z1 1 + = h(y)m(dy) ∞ z log+ zP (X (φh) dz) λ σ y mσ ∈ − 1 ZE Z1 1 = E X (φh) log+ X (φh) h(y)m(dy). (5.19) λ σ y mσ mσ − 1 ZE   23 The last term is finite by Lemma 5.1. Thus we get

+ + ∞ ∞ E [S S ]= E E (S ) E (S ) < + . x mσ,nσ − mσ,nσ x x mσ,nσ|Fnσ − x mσ,nσ|Fnσ ∞ nX=1 nX=1 h i e e Recall that S S for every n 0. Again by the Borel-Cantelli lemma, we have mσ,nσ ≥ mσ,nσ ≥

lim Smσ,nσ eSmσ,nσ = lim Ex(Smσ,nσ nσ) Ex(Smσ,nσ nσ) = 0 Px-a.s.. (5.20) n + n + → ∞ − → ∞ |F − |F Now (5.10) follows frome (5.15) and (5.20). e Substituting nσ by t and mσ by s in (5.13), we get for any t> 1,

2 λ1t λh/2 1/2 λht λ1s Ex Ss,t Ex Ss,t t e h(x) gs,t, h + e h(x)a1(x) e− − φ . − |F ≤ h i k k∞     e e + λ1te Using a similar calculation as in (5.14), we get for s> 0, 1 ∞ e gs,t, h dt < + . Consequently 2 h i ∞ λ1t limt + e gs,t, h = 0. Hence Ex Ss,t Ex Ss,t t R 0 as t + , or equivalently → ∞ h i − |F → → ∞     e e 2 lim Ss,t Ex(Ss,t t) =0 in L (Px). (5.21) t + − |F → ∞   Substituting nσ by t and mσ by s ine (5.18), wee get for any t> 1,

E S S = E E (S ) E S x s,t − s,t x x s,t|Ft − x s,t|Ft h i h λh/2  1/2 i λht λ1s h(x) fs,t, h + e a1(x) h(x)e− − φ . e ≤ h i e k k∞ + By similar calculation as in (5.19), we get ∞ fs,t, h dte < + for all s> 0. Hence limt + fs,t, h = 0 h i ∞ → ∞h i 0. Thus we have R

E S S = E E (S ) E S 0, as t + . (5.22) x s,t − s,t x x s,t|Ft − x s,t|Ft → → ∞ h i h  i Therefore (5.9) follows frome (5.21) and (5.22). e

Lemma 5.3. Under the conditions of Theorem 1.3, we have

λ1(s+t) 1 lim lim e Ex [Xs+t(φh) t]= M φh, h in L (Px), (5.23) s + t + ∞ → ∞ → ∞ |F h i for every φ +(E) and every x E. ∈Bb ∈ 1 Proof. Recall that Mt converges to M in L (Px) by Theorem 1.2. It suffices to prove that ∞

λ1(s+t) lim lim Ex e Ex(Xs+t(φh) t) Mt φh, h = 0. (5.24) s + t + |F − h i → ∞ → ∞ h i

Note that by the Markov property,

eλ1(s+t)E (X (φh) ) = eλ1(s+t) E [X (φh)] x s+t |Ft Xu(t) s u t X∈Z

24 λ1t h = e Xt hPs (φ) . (5.25)   Thus we have for any s,t>t0,

E eλ1(s+t)E (X (φh) ) M φh, h x x s+t |Ft − th i h i eλ1tE X hP hφ h φh, h ≤ x t s − h i h   i eλ1tE X hP hφ h φh, h ≤ x t s − h i h  i = h(x)P h P h φ φh, h (x) . t s − h i  

Using (2.8) we continue the estimation above to get:

E eλ1(s+t)E (X (φh) ) M φh, h x x s+t |Ft − th i λhh (s t0/2) h 1/2 i e− − φ h(x)Pt (at )(x) ≤ k k∞ 0 1/2 λh(s t0/2) 1/2 1/2 λh(t t0/2) e− − φ h(x) eat0 (y) m(dy)+ at0 (x) e− − at0 (y)m(dy) ≤ k k∞ "ZE ZE  # 0 as t + , and then s + . → → ∞ e → ∞e e e e

Proof of Theorem 1.3: Let φ = f/h, then φ +(E). Note that for any s,t > 0 ∈Bb

λ1(t+s) λ1(t+s) λ1(s+t) e Xt+s(φh) M φh, h e Xt+s(φh) e Ex (Xs+t(φh) t) − ∞h i ≤ − |F λ1(s+t) + e Ex (Xs+t(φh) t) M φh, h . |F − ∞h i

Therefore Theorem 1.3 follows immediately from Lemma 5.2 and Lemma 5.3.

6 Strong law of large numbers

In this section, we prove Theorem 1.5.

6.1 SLLN along lattice times

Lemma 6.1. Suppose the assumptions of Theorem 1.5 hold. Then for any σ > 0 and any x E, ∈

λ1nσ lim e Xnσ(φh)= M φh, h Px-a.s. (6.1) n + ∞ → ∞ h i for every φ +(E) ∈Bb

λ1nσ Proof. Recall that M = limn + e Xnσ(h). Let g(y) := (φh)(y) h(y) φh, h . The conver- ∞ → ∞ − h i gence in (6.1) is equivalent to

λ1nσ lim e Xnσ(g) = 0 Px-a.s. (6.2) n + → ∞

25 For an arbitrary m N, ∈

λ1(n+m)σ e X(n+m)σ(g) = eλ1(n+m)σX (g) E eλ1(n+m)σX (g) + E eλ1(n+m)σX (g) (n+m)σ − x (n+m)σ Fnσ x (n+m)σ Fnσ  h i h i =: In + IIn. (6.3)

Note that by Lemma 5.2, we have

I = eλ1(n+m)σX (φh) E eλ1(n+m)σX (φh) 0 as n + P -a.s. (6.4) n (n+m)σ − x (n+m)σ Fnσ → ↑ ∞ x h i

On the other hand, by Markov property, we have

λ1(n+m)σ IIn = e EXu(nσ)(Xmσ(g)) u nσ ∈ZX λ1(n+m)σ (Q 1)µ = e Pmσ− g(Xu(nσ)) u nσ ∈ZX λ1(n+m)σ (Q 1)µ = e Xnσ(Pmσ− g). (6.5)

Note that our assumption (1.14) implies (1.17). Then for any fixed ε > 0, there exist m > 0 sufficiently large such that sup ph(mσ,x,y) 1 ε. x,y E | − |≤ ∈ Then for any x E, ∈

λ1mσ (Q 1)µ h e P − g(x) = h(x)P (g/h)(x) | mσ | | mσ | = h(x) ph(mσ,x,y) 1 φ(y)m(dy) − ZE   h 2 h(x) p (mσ,x,y) 1 φ(y)h(ye) m( dy) ≤ E − Z εh(x) φh, h . ≤ h i Consequently by (6.5), we have II ε φh, h M . (6.6) | n|≤ h i nσ It follows from (6.3), (6.4) and (6.6) that

λ1nσ lim sup e Xnσ(g) ε φh, h M Px-a.s.. n + | |≤ h i ∞ → ∞ Hence we get (6.2) by letting ε 0. → 6.2 Transition from lattice times to continuous time

In this subsection we extend the convergence along lattice times in Lemma 6.1 to convergence along continuous time and give a sketch of proof of Theorem 1.5. The main approach in this subsection is similar to that of [6, Theorem 3.7] (see also [3, Theorem 1’]). According to the proof of [6, Theorem 3.7], to prove Theorem 1.5, it suffices to prove the following lemma:

26 Lemma 6.2. Under the conditions of Theorem 1.5, for every open subset U in E and every x E, ∈

λ1t 2 lim inf e Xt(1U h) M h(y) m(dy) Px-a.s., (6.7) t + ≥ ∞ → ∞ ZU Note that if (6.7) is true for any bounded open set U in E with U E, then (6.7) is true for ⊂ any open set U. In fact, for an arbitrary open set U in E, there exists a sequence of bounded open sets U : n 0 such that U E and U U. Hence if U satisfies (6.7), we can deduce that U { n ≥ } n ⊂ n ↑ n satisfies (6.7) by monotone convergence theorem. In the following we assume that U is an arbitrary bounded open set in E with U E. For any ⊂ ε,σ > 0, n N and x E, define ∈ ∈ 1 U ε(x) := y U : h(y) h(x) , { ∈ ≥ 1+ ε }

σ,ε 1 Y := (h1 )(X (nσ))1 ε u,nσ , n,u U u Xv(t) U (Xu(nσ)) for all v and t [nσ,(n+1)σ] 1+ ε { ∈ ∈Lt ∈ } and σ,ε λ1nσ σ,ε Sn := e Yn,u . u nσ ∈ZX Here for each t nσ, u,nσ denotes the set of particles which are alive at t and are descendants of ≥ Lt the particle u . ∈Lnσ To prove Lemma 6.2, we need the following lemma.

Lemma 6.3. Suppose that U is an arbitrary bounded open set U in E with U E. Under the ⊂ conditions of Theorem 1.5, we have

σ,ε σ,ε lim S Ex [S nσ] = 0 Px-a.s. (6.8) n + n n → ∞ − |F for every x E. ∈ Proof. Note that for n N and Y : i = 1, ,n independent real-valued centered random ∈ { i · · · } variables, n n n E Y 2 = Var Y = E Y 2. | i| i | i| Xi=1 Xi=1 Xi=1 Thus by the conditional independence between subtrees and the Markov property of a branching Markov process, we have

2 E (Sσ,ε E [Sσ,ε ])2 = e2λ1nσ E Y σ,ε E Y σ,ε x n − x n |Fnσ Fnσ x n,u − x n,u |Fnσ Fnσ h i u nσ h i ∈ZX   2e2λ1nσ E Y σ,ε 2 ≤ x | n,u | |Fnσ u nσ ∈ZX   2λ1nσ E σ,ε 2 = 2e Xu(nσ) Y0, | ∅ | u nσ ∈ZX h i 2 2λ1nσ 2 2(1 + ε)− e (h 1 )(X (nσ)). ≤ U u u nσ ∈ZX 27 Thus

+ ∞ ∞ σ,ε σ,ε 2 2 λ12nσ 2 E (S E [S ]) 2(1 + ε)− e E (h 1 )(X (nσ)) x n − x n |Fnσ ≤ x U u n=1 n=1 "u nσ # X h i X ∈ZX ∞ 2 λ12nσ (Q 1)µ 2 2(1 + ε)− e P − (h 1 )(x) ≤ nσ U n=1 X ∞ 2 λ12nσ h 2(1 + ε)− h(x) e P (h1 )(x). ≤ nσ U nX=1 By Proposition 1.8, for any ε > 0, ph(nσ,x,y) 1 ε for every x,y E when n is sufficiently | − | ≤ ∈ large. It follows that P h (h1 )(x) (1 + ε) h(y)m(dy) < + , and so nσ U ≤ U ∞ + R ∞ ∞ E (Sσ,ε E [Sσ,ε ])2 e ch(x) eλ1nσ < , x n − x n |Fnσ ≤ ∞ n=1 n=1 X h i X where c is a positive constant. This together with the Borel-Cantelli lemma yields (6.8).

Proof of Lemma 6.2 : If U is an arbitrary bounded open set in E with U E, then (6.7) follows ⊂ from Lemma 6.3 in the same way as [6, Theorem3.9]. We omit the details here. By the argument after (6.7), we know that (6.7) holds for any open subset U of E.

Acknowledgement The research of Zhen-Qing Chen is partially supported by NSF grant DMS- 1206276 and NNSFC 11128101. The research of Yan-Xia Ren is supported by NNSFC (Grant No. 11271030 and 11128101). The research of Ting Yang is partially supported by NNSF of China (Grant No. 11501029) and Beijing Institute of Technology Research Fund Program for Young Scholars.

References

[1] Albeverio, S., Blanchard, P. and Ma, Z.-M.: Feynman-Kac semigroups in terms of signed smooth measures, in “Random Partial Differential Equations”, (U.Hornung et al. Eds) Birkhauser, Basel, 1–31, (1991).

[2] Albeverio, S. and Ma, Z.-M.: Perturbation of Dirichlet forms-lower semiboundedness, closability, and form cores, J. Funct. Anal. , 99, 332–356, (1991).

[3] Asmussen, S. and Hering, H.: Strong limit theorems for general supercritical branching processes with applications to branching diffusions. Z. Wahrsch. Verw. Gebiete, 36, 195–212, (1976).

[4] Chen, Z.-Q., Fitzsimmons, P. J., Takeda, M., Ying, J. and Zhang, T.-S.: Absolute continuity of sym- metric Markov processes, Ann. Probab., 32, 2067–2098, (2004).

[5] Chen, Z.-Q. and Fukushima, M.: Symmetric Markov Processes, Time Changes and Boundary Theory, Princeton University Press, USA, (2012).

[6] Chen, Z.-Q. and Shiozawa, Y.: Limit theorems for branching Markov processes, J. Funct. Anal., 250, 374–399, (2007).

28 [7] Dunford, N. and Schwartz, J.: Linear Operator, vol. I, Interscience, New York, (1958). [8] Durrett, R.: and Examples, 4nd edn., Cambridge University Press, (2010). [9] Eckhoff, M., Kyprianou, A. E. and Winkel, M.: Spine, skeletons and the strong law of large numbers. Ann. Probab., to appear, (2015). [10] Engl¨ander, J.: Law of large numbers for superdiffusions: The non-ergodic case. Ann. Inst. Henri Poincar´eProbab. Stat., 45, 1–6, (2009). [11] Engl¨ander, J., Harris, S. C. and Kyprianou, A. E.: Strong law of large numbers for branching diffusions. Ann. Inst. Henri Poincar´eProbab. Stat., 46, 279–298, (2010). [12] Engl¨ander, J. and Kyprianou, A. E.: Local extinction versus local exponential growth for spatial branch- ing processes. Ann. Probab., 32, 78–99, (2003). [13] Fukushima, M., Oshima, Y. and Takeda, M.: Dirichlet Forms and Symmetric Markov Processes, Walter de Gruyter, Berlin, (1994). [14] Hardy, R. and Harris, S. C.: A spine approach to branching diffusions with applications to Lp- convergence of martingales. In S´eminaire de Probabiliti´es XLII. C. Donad-Martin, M. Emery.´ A. Rouault and C. Stricker (Eds), 1979, 281–330, (2009). [15] Kaleta, K., L˝orinczi, J.: Analytic properties of fractional Schr¨odinger semigroups and Gibbs measures for symmetric stable processes, arXiv:1011.2713, preprint, (2010). [16] Kim, P. and Song, R.: Estimates on Green functions and Schr¨odinger-type equations for non-symmetric diffusions with measure-valued drifts, J. Math. Anal. Appl., 332, 57–80, (2007). [17] Kim, P. and Song, R.: Two-sided estimates on the density of Brownian motion with sigular drift, Illinois J. Math., 50(3), 635–688, (2006). [18] Liu, R.-L. Ren, Y.-X. and Song, R.: L log L criteria for a class of superdiffusons. J. Appl. Probab., 46, 479–496, (2009). [19] Liu, R.-L. Ren, Y.-X. and Song, R.: L log L conditions for supercritical branching Hunt processes, J. Theor. Probab., 24, 170–193, (2011). [20] Ren, Y.-X., Song, R. and Zhang, R.: Central limit theorems for supercritical branching Markov pro- cesses, J. Funct. Anal., 266, 1716–1756, (2014). [21] Riahi, L.: Comparison of Green Functions and Harmonic Measures for Parabolic Operators. Potential Analysis, 23, 381–402, (2005). [22] Schaeffer, H. H.: Banach Lattices and Positive Operators, Springer, New York, (1974). [23] Shiozawa, Y.: Exponential growth of the numbers of particles for branching symmetric α-stable pro- cesses. J. Math. Soc. Japan, 60, 75–116, (2008). [24] Stollman, P. and Voigt, J.: Perturbation of Dirichlet forms by measures, Potential Anal., 5, 109-138, (1996). [25] Watanabe, S.: On the branching process for Brownian particles with an absorbing boundary, J. Math. Kyoto Univ., 4, 385–398, (1965). [26] Watanabe, S.: Limit theorems for a class of branching processes, in “Markov Processes and Potential Theory” (J. Chover (Eds.), Wiley, New York, 205–232, (1967).

29