Skip to content
BY 4.0 license Open Access Published by De Gruyter February 25, 2022

Continuous flows driving branching processes and their nonlinear evolution equations

  • Lucian Beznea EMAIL logo and Cătălin Ioan Vrabie

Abstract

We consider on M(ℝd) (the set of all finite measures on ℝd) the evolution equation associated with the nonlinear operator FΔF+k1bkFk , where F′ is the variational derivative of F and we show that it has a solution represented by means of the distribution of the d-dimensional Brownian motion and the non-local branching process on the finite configurations of M(ℝd), induced by the sequence (bk)k⩾1 of positive numbers such that k1bk1 . It turns out that the representation also holds with the same branching process for the solution to the equation obtained replacing the Laplace operator by the generator of a Markov process on ℝd instead of the d-dimensional Brownian motion; more general, we can take the generator of a right Markov process on a Lusin topological space. We first investigate continuous flows driving branching processes. We show that if the branching mechanism of a superprocess is independent of the spatial variable, then the superprocess is obtained by introducing the branching in the time evolution of the right continuous flow on measures, canonically induced by a right continuous flow as spatial motion. A corresponding result holds for non-local branching processes on the set of all finite configurations of the state space of the spatial motion.

MSC 2010: 35J60; 60J35; 60J80; 60J68; 60J45; 47D07

1 Introduction

Let (bk)k⩾1 be a sequence of positive numbers such that k1bk1 and consider the following parabolic equation on measures

(1.1) ddtFt=ΔFtcFt+ck1bkF tkonM(d),t0,

with the initial condition F0 = F, where c > 0 and for a function F : M(ℝd) → [0, 1] we denoted by F′ its variational derivative.

The first goal of this paper is to show that equation (1.1) has a mild solution having a probabilistic representation by means of the distribution of the d-dimensional Brownian motion and the non-local pure branching process on the finite configurations of M(ℝd), induced by the sequence (bk)k⩾1. Recall that the measure-valued branching processes intend to be stochastic models for the population dynamics. The time evolution of a system of particles is characterised by a branching mechanism and a spatial motion which describes the movement of the particles between the branching moments. The spatial motion is a right Markov space with state space a Lusin topological space E, while the state space of the branching process is a set of measures on E, namely, either M(E) (:= the set of all finite measures on E) for the superprocesses (see [1, 2, 15, 16, 18, 19, 22, 29, 30]), or Ê (:= the set of all finite configurations of E) for the non-local branching processes; see e.g. [8, 10, 11, 26, 31].

If in (1.1) we replace the Laplace operator with a second order integro-differential operator generating a Markov process with state space ℝd, then the obtained equation has a solution with a similar probabilistic representation, and notice that it is based on the same pure branching process. In this way we complete the answer given in [8] to the challenging suggestion of T.E. Harris from [23] , page 50, to investigate branching processes for which “each object is a very complicated entity; e.g., an object may itself be a population”. In fact, in Section 5 we consider equation (1.1) on M(E), with the generator of a right Markov process on the Lusin topological space E instead of the Laplacian on ℝd.

Our second goal is to emphasize a class of branching processes which are driven by a right continuous flow Φ on E, in the sense that such a measure-valued process X^ admits a representation by means of a second branching process X0^ and of the flow on measures induced by Φ,

(1.2) X^t=Φt(X t0^)forallt0.

It turns out that the branching process X0^ has the same branching mechanism as X^ , however X0^ has no spatial motion and therefore we call it pure branching process.

The representation (1.2) holds in the case of a superprocess for which the branching mechanism Ψ is independent of the spatial variable, Ψ : [0, ∞) → ℝ is of the form

Ψ(λ)=bλcλ2+0(1eλsλs)N(s˙),

where b, c ∈ ℝ, c ⩾ 0, and N is a measure on (0, ∞) such that N(uu2) < ∞, and the spatial motion is a right continuous flow on E, regarded as a deterministic Markov process. Recall that a superprocess X^ provides a stochastic solution for the nonlinear evolution equation written formally as

(1.3) ddtvt=Dvt+Ψ(vt),t0,

with the initial condition v0 = f, where D is the generator of the spatial motion and Ψ is the branching mechanism. The nonlinear evolution equation associated with the pure branching process X0^ is the particular case of (1.3) with no spatial motion term on the right hand side, that is, with D ≡ 0. Actually, we consider the integral version of equation (1.3), or equivalently, we work with the mild solutions of it.

A basic step of our approach is to solve equation (1.3) with D the generator of a right continuous flow Φ on E, to find a solution vt, t ⩾ 0, having the following two probabilistic representations:

(1.4) vt(x)=ln𝔼0^δxefΦt(X t0^)=ln𝔼0^δΦt(x)ef(X t0^),xE,t0,

where ef is the exponential mapping on M(E), ef (μ):=e−∫fdμ for all μM(E); with the terminology of [18] , page 133, vt is written as the “log-potential of X t0^ and Φt”. We prove analogous results for non-local branching processes, provided that the flow and the distributions of the branching are compatible.

If 𝒧 and 𝒧0 are the extended weak generators of the superprocesses X^ and respectively X0^ on M(E), then we have

𝒧=D^+𝒧0,

where D^ is the generator of the continuous flow on M(E) induced by Φ. In this way, the representation (1.2) may be interpreted as a consequence of regarding 𝒧 as a modification of 𝒧0 with the first order operator D^ , which is a substitute for a “drift type” operator acting in the considered infinite dimensional frame. This is an exemplification of a general strategy developed in [3] ; for more details see assertion (iv) of Remark 3.5 below.

A key point in obtaining the claimed probabilistic solution to (1.1) is the representation (1.2) applied on M(ℝd) as state space for the spatial motion.

We present now the structure of the paper.

The first main result (Theorem 3.2) is stated in Section 3. It is preceded by the description of the weak generator of a superprocess (Proposition 3.1), which completes results from [2, 19, 22, 30]. Results related to equation (1.3) and (1.4) are gathered in assertion (iii) of Theorem 3.2. Corollary 3.4, and Remark 3.5.

The non-local branching processes on configuration spaces are investigated in Section 4. First, we highlight relevant absorbing sets of finite configurations with respect to a pure branching process on Ê (in Proposition 4.1). Theorem 4.2 is the second main statement of the work. Results similar to (1.4) and Corollary 3.4, but valid on Ê, are collected in Remark 4.3, which also indicates absorbing sets for the pure branching process, when the branching mechanism is the classical one.

The proofs of the two theorems use a unifying argument (Lemma 3.3), which is applicable on both spaces of measures (M(E) and Ê) and we put its proof in the Appendix. Recall that in [9] it was developed a method of proving the path regularity for both superprocesses and non-local branching processes. The Appendix also presents briefly complements on the right continuous flows and the extended weak generator of a Markov process.

The application in solving equation (1.1) is given in Section 5, Corollary 5.2.

In Section 2 we present basic notations and preliminary facts on the right Markov processes, branching processes, and potential theoretical tools we shall use further: transition semigroups and resolvents of kernels, excessive functions, the fine topology, absorbing sets, and right continuous flows on measures.

2 Preliminaries

For a σ-algebra 𝒢 we denote by [𝒢] (resp. p𝒢) the vector space of all real-valued (resp. the set of all positive, numerical) 𝒢-measurable functions on E. Also, for a set of real-valued functions 𝒞 we denote by σ(𝒞) the σ-algebra generated by 𝒞, by [𝒞] the vector space spanned by 𝒞, and by p𝒞 (resp. b𝒞) the set of all positive (resp. bounded) functions from 𝒞.

Excessive functions. Let E be a Lusin topological space (i.e., E is homeomorphic to a Borel subset of a compact metric space) with Borel σ-algebra 𝒝(E). Let further 𝒰 = (Uα)α>0 be a sub-Markovian resolvent of kernels on (E, 𝒝(E)). We denote by 𝒠(𝒰) the set of all 𝒝(E)-measurable 𝒰-excessive functions, i.e., u𝒠(𝒰) if and only if u is a nonnegative, numerical, 𝒝(E)-measurable function, αUαuu for all α > 0, and limα→∞ αUαu(x) = u(x), xE.

If β > 0, we denote by 𝒰β the sub-Markovian resolvent of kernels 𝒰β = (Uβ+α)α>0. A 𝒰β-excessive function is also called β-excessive.

Let X = (Ω, 𝒡, 𝒡t, Xt, θt, 𝕇x, ζ) be the right Markov process on E having 𝕋 = (Tt)t⩾0 as transition function, Ttf(x) = 𝔼x(f(Xt), t < ζ), t ⩾ 0, f ∈ p𝒝(E) (:= the set of all positive 𝒝(E)-measurable functions on E). The resolvent of kernels 𝒰 = (Uα)α>0 on (E, 𝒝(E)) associated with (Tt)t⩾0 is defined as Uαf:=0eαtTtfdt , f ∈ p𝒝(E).

Fine topology. The fine topology is the coarsest topology on E making continuous all β-excessive functions for some (and equivalently for all) β > 0. Recall that a function f from p𝒝(E) is finely continuous if and only if tf(Xt) is a.s. right continuous on [0, ζ). Consequently, due to the right continuity of the paths of the process X, each continuous function on E is finely continuous; see [4, 32] for details on the potential theory for Markov processes with a general state space.

Absorbing set. A set A𝒝(E) is called absorbing (with respect to the Markov process X) if R βE\A1=0 on A, where R βA1:=inf{u𝒠(𝒰β):u1A} for some β > 0. If A is absorbing then it is finely open, Uβ1E\A = 0 on A, and the restriction of X to A is again a right process; see e.g. [32] or [13] . Basic examples of absorbing sets are [v = 0] and [v < ∞] for some v𝒠(𝒰β). Examples of absorbing sets for branching processes will be given in Proposition 4.1 and Remark 4.3 (iii) below.

Trivial Markov process. Let X0=(X t0)t0 be the Markov process on E for which each point is a trap, i.e., 𝕇x(X t0=x)=1 for all t ⩾ 0 and xE, or equivalently, each kernel from its transition function (T t0)t0 is the identity operator, T t0f=f for every t ⩾ 0 and f ∈ p𝒝(E). Clearly, any Borel subset of E is absorbing with respect to the trivial Markov process on E.

Branching processes. We endow M(E) with the weak topology and let 𝒨(E) denote the Borel σ-algebra on M(E).

A second measure state space we will consider is the set ÊM(E) of all finite configurations of E, E^:={ ii0δxi:i0𝕅,i01,xiEforall1ii0 }{0} , where 0 denotes the zero measure. The set Ê is endowed with the weak topology on the finite measures on E and the corresponding Borel σ-algebra 𝒝(Ê); see e.g. [14] .

Further (M, 𝒨) will denote either (M(E), 𝒨(E)) or (Ê, 𝒝(Ê)). Recall that a right Markov process with state space M is called branching process provided that for any two independent copies X and X′ of the given process on M, starting respectively from two measures μ and μ′, X + X′ and the process starting from μ + μ′ are equal in distribution. Recall also that if p1, p2 are two finite measures on M, then their convolution p1 * p2 is the finite measure on M defined for every h ∈ p𝒨 by

Mp1*p2(dν)h(ν):=Mp1(dν1)Mp2(dν2)h(ν1+ν2).

A bounded kernel Q on (M, 𝒨) is called branching kernel provided that

Qμ+ν=Qμ*Qνforallμ,νM,

where Qμ denotes the measure on M such that MhdQμ=Qh(μ) for all h ∈ p𝒨. It turns out that: a right Markov process with state space M is a branching process if and only if its transition function is formed by branching kernels. For a function f ∈ bp𝒝(E) we shall consider the mappings lf : M → ℝ+ and ef : M → [0, 1] defined as

lf(μ):=μ,f:=Efdμ,μM,ef:=elf.

Notice that 𝒨 is generated by {lf : f ∈ bp𝒝(E)} and also by {ef : f ∈ bp𝒝(E)}.

Right continuous flows on measures. We present in Appendix (A3) the basic facts on the right continuous flows.

If t ⩾ 0 define the kernel Q t0 on M(E) as

Q t0F(μ):=F(μTt),Fp𝒨(E),μM(E).

We may consider the right continuous flow on M(E) induced by the right Markov process X. More precisely, let Φ^=(Φ^t)t0 be the family of mappings on M(E) defined as Φ^t(μ):=Xt(𝕇μ) , μM(E); recall the notation Xt(𝕇μ)(A):=E𝕇x(X t1(A))μ(dx) for all A𝒝(E). Then Φ^ is a right continuous flow on M(E) and its transition function is 𝕈0=(Qt0)t0 . It is easy to check that Q t0 , t ⩾ 0, is a branching kernel on M(E) and consequently, the (deterministic) Markov process induced by Φ^ is actually a branching process on M(E), it is the (X, 0)-superprocess.

Let Φ = (Φt)t⩾0 be a right continuous flow on E. Recall that Φ may be regarded as a deterministic right Markov process on E with infinite lifetime, X = (Ω, 𝒡, 𝒡t, Xt, θt, 𝕇x): Ω = E, 𝒡 = 𝒡t = 𝒝(E), Xt(x) := Φt(x) =: θt(x) for all xΩ, and 𝕇x = δx. In this particular case, the right continuous flow on M(E) induced by X, as described before, is also denoted by Φ = (Φt)t⩾0: if μM(E) then Φt(μ) is the transport of μ by means of Φt, i.e.,

Φt(μ)(A):=μ(Φ t1(A))forallA𝒝(E)andt0.

If μÊ, μ=i=1kδxi , then Φt(μ) also belongs to Ê, Φt(i=1kδxi)=i=1kδΦt(xi) , t ⩾ 0. Consequently, the right continuous flow Φ on E also induces a right continuous flow on Ê, we shall denote it by Φ too; here we put Φt(0) = 0.

Let A𝒝(E) be an absorbing set with respect to the right continuous flow Φ on E, i.e., Φt(x) ∈A for all xA. Then  is an absorbing set with respect to the right continuous flow on Ê induced by Φ.

3 Right continuous flows driving superprocesses

In the sequel the fixed right Markov process X = (Ω, 𝒡, 𝒡t, Xt, θt, 𝕇x) with state space E and infinite lifetime is called spatial motion, and let 𝒰 be its resolvent. We also fix a branching mechanism, that is, a function Ψ : E × [0, ∞) → ℝ of the form

Ψ(x,λ)=b(x)λc(x)λ2+0(1eλsλs)N(x,ds),

where c ⩾ 0 and b are bounded 𝒝(E)-measurable functions, and N : p𝒝((0, ∞)) → p𝒝(E) is a kernel such that N(uu2) ∈bp𝒝(E). Examples of branching mechanisms are Ψ(λ) = −λα for 1 < α ⩽ 2 and Ψ(λ) = λα for 0 < α < 1.

We present now briefly the construction of the measure-valued branching Markov process associated with X and Ψ, the (X, Ψ)-superprocess; cf. [2, 22, 30].

Equation (1.3) has a unique mild solution, more precisely, for each f ∈ bp𝒝(E) the equation

(3.1) vt(x)=Ttf(x)+0tTs(x,Ψ(,vts))ds,t0,xE,

has a unique solution (t, x) ↦ Vtf(x) jointly measurable in (t, x) such that sup0⩽st ||Vsf|| < ∞ for all t > 0 and [0, ∞) ∋ tVtf(x) is right continuous for all xE, provided that tTtf(x) has this property.

The mappings fVtf form a nonlinear semigroup of operators on bp𝒝(E).

For each t ⩾ 0 there exists a unique Markovian kernel T^t on (M(E), 𝒨(E)) such that

(3.2) T^t(ef)=eVtf,fbp𝒝(E).

Since the family (Vt)t⩾0 is a (nonlinear) semigroup on bp𝒝(E), it follows that 𝕋^=(𝕋^t)t0 is a transition function (M(E), 𝒨(E)). By Theorem 4.9 from [2] , under a Feller-type regularity condition, there exists a Borel right process X^ , with state space M(E), having the transition function T^ , called (X, Ψ)-superprocess.

Let 𝒰^=(U^α)α>0 be the Markovian resolvent of kernels on (M(E), 𝒨(E)) induced by 𝕋^ .

Let

β:=||b||andb:=b+βwithββ.

Then b′ ⩾ 0 and consider the transition function 𝕋b=(T tb)t0 of the process obtained by killing X with the multiplicative functional induced by b′,

T tbf(x)=𝔼x(eotb(Xs)dsf(Xt)),fp𝒝(E),xE.

If u ∈ bp𝒝(E) then by Corollary 4.3 in [2] we have

(3.3) u𝒠(𝕋b)lu𝒠(𝕋^β)1eαu𝒠(𝕋^β)forallα>0.

We also have

(3.4) T^t(lu)=lT tbu.

The extended weak generator of a superprocess. Let (L, 𝒟(L)) be the extended weak generator of the spatial motion X on E and (𝒧, 𝒟(𝒧)) the extended weak generator of the (X, Ψ)-superprocess on M(E); see Appendix (A4) for the basic notions about the extended weak generator of a Markov process.

Let (D^,𝒟(D^)) be the extended weak generator of the right continuous flow Φ^ on M(E) induced by the right Markov process X, Φ^t(μ):=Xt(𝕇μ) , μM(E), that is, the extended weak generator of the transition function 𝕈0=(Q t0)t0 , Q t0F(μ):=F(μTt) for all F ∈ bp𝒨(E), μM(E), and t ⩾ 0.

Proposition 3.1

  1. If ub𝒟 oc(L) is such that Lu is a bounded function then lu belongs to the domain 𝒟(𝒧) of the extended weak generator (𝒧, 𝒟(𝒧)) of X^ and

    𝒧(lu)=lLubu.

  2. Let n ⩾ 1, u1, . . . , un ∈ b𝒟(L) such that all Lui are bounded functions, ψC b1(n) , and consider the function F on M(E), defined as F(μ) := ψ(〈μ, u1〉, . . . , 〈μ, un〉), μM(E). Then F belongs to 𝒟(D^) , F′(μ, ·) ∈ 𝒟(L) and we have

    (3.5) D^F(μ):=ELF(μ,x)μ(dx)forallμM(E),

    where recall that the variational derivative of a function F : M(E) → ℝ is F(μ,x):=limɛ01ɛ(F(μ+ɛδx)F(μ)) , μM(E), xE.

We present the proof of Proposition 3.1 in Appendix (A1). Notice that assertion (ii) is a particular case of a version of a result from [22] ; see also Remark 3.5 (i) below.

Further, we consider branching processes with no spatial motion.

Pure branching superprocesses. The (X′, Ψ)-superprocess is named pure branching. It is a right Markov process X0^ with state space M(E) and its transition function (T t0^)t0 is obtained as follows. The particular case of the equation (3.1),

(3.6) vt(x)=f(x)+0tΨ(x,vs(x))ds,t0,xE,

has a unique solution (t,x)V t0f(x) jointly measurable in (t, x). For each t ⩾ 0 there exists a unique Markovian kernel T t0^ on (M(E), 𝒨(E)) such that T t0^(ef)=eV t0f , f ∈ bp𝒝(E). So, the time evolution of the process X0^ is governed only by the branching procedure.

For a nonlinear evolution equation which is equivalent to (3.6) see Remark 3.5 below.

We can state now the first main result of this paper, involving the pure branching process and the right continuous flow driving it. In particular, this is an illustration of Section 4 from [3] .

Theorem 3.2

Consider a branching mechanism Ψ which is independent of the spatial variable, that is, with b and c constant functions and N(x, ds) = N(ds) for all xE. Assume that the spatial motion X = (Xt)t⩾0 is the deterministic process associated with a right continuous flow Φ = (Φt)t⩾0 on E and suppose that the mapping [0, ∞) × E ∋ (t, x) ↦ Φt(x) is continuous.

Let X^=(X^t)t0 be the (X, Ψ)-superprocess induced by the spatial motion X and the branching mechanism Ψ and let X0^=(X t0^)t0 be the pure branching (X0 , Ψ)-superprocess. Then the following assertions hold.

  1. The (X, Ψ)-superprocess X^=(X^t,𝕇^μ) has the representation (1.2) as the pure branching process X0^=(X t0^,𝕇0^μ) driven by the continuous flow Φ on M(E), where the equality is in the distribution sense and we also have

    (3.7) 𝕇^μ(X^tΓ)=𝕇^0^Φt(μ)(X t0^Γ)forallμM(E)andΓ𝒨(E).

  2. Let (D, 𝒟(D)) (resp. (D^,𝒟(D^)) , (𝒧, 𝒟(𝒧)), and (𝒧0 , 𝒟(𝒧0))) be the extended weak generator of the spatial motion X (resp. of the flow Φ on M(E), X^ , and X0^ ). Let further 𝒟o := UαVβ(bC(M(E))), α, β > 0, where (Uα)α>0 (resp. (Vα)α>0) is the resolvent of X0^ (resp. the resolvent of the flow Φ on M(E)). Then 𝒟o𝒟 oc(𝒧)𝒟o(D^)𝒟(𝒧0) and

    𝒧=D^+𝒧0on𝒟o.

  3. Equation (1.3) has a mild solution vt, t ⩾ 0, which is given by the log-potential type formula (1.4). More precisely, vt is the solution of the nonlinear integral evolution equation (3.1).

Proof

We use the construction of V t0f from [2] , the proof of Proposition 3.2. Observe that, using Proposition 3.1 from [2] , equation (3.6) is equivalent to

(3.8) vt(x)=eatf(x)+0tea(ts)Ψ0(vs(x))s˙,t0,xE

where Ψ0 (λ) := Ψ(λ) + , and let V t0f , t ⩾ 0, be its unique solution for f ∈ bp𝒝(E).

We need the following lemma, we present its proof in Appendix (A2).

Lemma 3.3

Let J : EE be a 𝒝(E)/𝒝(E)-measurable map and denote by 𝒦(J) the set of all operators (W, 𝒟(W)) on bp𝒝(E), W : 𝒟(W) → bp𝒝(E), such that for all f ∈ bp𝒟(W) we have f ○ J𝒟(W) and W(f ○ J) = WfJ. Convention: If the domain 𝒟(W) of W is not indicated, then this means that 𝒟(W) = bp𝒝(E).

Then the following assertions hold.

  1. The set 𝒦(J) has the following properties.

    1. If (W, 𝒟(W)) is an operator on bp𝒝(E) such that there exists a sequence (Wn)n in 𝒦(J) with 𝒟(Wn) = 𝒟(W) for all n, which is converging pointwise to W (i.e. limn Wnf(x) = Wf(x) for all f𝒟(W) and xE), then W belongs to 𝒦(J);

    2. If (W, 𝒟(W)) and (V, 𝒟(V)) are two operators on bp𝒝(E) such that W(𝒟(W)) ⊂ 𝒟(V) and V, W𝒦(J) then VW𝒦(J);

    3. Let (K, 𝒟(K)) be an operator on bp𝒝(E) such that K𝒦(J). If Wf=0tK(Wsf)ν(ds) , where ν is a finite measure on [0, t] and Ws𝒦(J), 𝒟(Ws) = 𝒟(K), and Ws(𝒟(Ws)) ⊂ 𝒟(K) for all s, then W also belongs to 𝒦(J).

  2. Assume that (K, 𝒟(K)) is Lipschitz with respect to the supremum norm and K(0) = 0. Let a > 0, to > 0 and consider the integral equation

    (3.9) wt=eatf+0tea(ts)K(ws)ds,0tto,

    where f𝒟(K).

  3. The equation (3.9) has a unique solution [0, to] ∋ tWtf𝒟(K) such that the function (t, x) ↦ Wtf(x) is jointly measurable, provided that one of the following two conditions is verified:

    1. 𝒟(K) = bp𝒝(E);

    2. 𝒟(K) = {f ∈ bp𝒝(E) : f ⩽ 1} and K fa for all f𝒟(K).

  4. Assume that K preserves the pointwise convergence, that is, if (fn)n𝒟(K) is pointwise converging to f𝒟(K), then K(fn) → K(f) pointwise on E. Then the function [0, to) ∋ tWtf(x) is differentiable for each xE and ut := Wtf is the unique solution of the nonlinear evolution equation

    (3.10) dutdt=aut+K(ut),0t<to,

    with the initial condition uo = f.

  5. Suppose in addition that K𝒦(J), then Wt𝒦(J) for all t ∈ [0, to].

Let J be a map as in Lemma 3.3. We claim that

(3.11) V t0𝒦(J)forallt0.

Assume first that c = 0 and 01N(ds)< . Let K f(x) := Ψ0 (f(x)) for all f ∈ bp𝒝(E) and xE. Then K is Lipschitz and one can see that K𝒦(J) and for t ∈ [0, to] equation (3.8) is precisely (3.9). We can apply Lemma 3.3 (iii) to conclude that V t0𝒦(J) .

In the general case for c and N, V t0 is obtained after an additional pointwise approximation and using again property (p1) we conclude that (3.11) holds.

Applying (3.11) for J = Φt′, t′ ⩾ 0, it follows that

(3.12) V t0(fΦt)=V t0fΦtforallfbp𝒝(E)andt,t0.

We can prove now assertion (i). Equality (1.2) is equivalent to

(3.13) T^tF(μ)=𝔼0^μF(Φt(X t0^))forallFbp𝒨(E),t0,andμM(E),

where recall that (T^t)t0 is the transition function of X^ . By a monotone class argument we may suppose that F = ef with f ∈ bp𝒝(E) and in this case the right hand side of the equality (3.13) equals T t0^(efΦt)(μ) , where (T t0^)t0 is the transition function of X0^ . By (3.2) applied to T^t and T t0^ it follows now that (3.13) is equivalent to

Vtf=V t0(fΦt)forallfbp𝒝(E)andt0.

Let (St)t⩾0 be the transition function of the right continuous flow Φ, Stf := fΦt for all f ∈ bp𝒝(E) and t ⩾ 0. So, by the uniqueness of the solution to equation (3.1), we have to check that (t,x)V t0(fΦt)(x) verifies (3.1), written with St instead of Tt. Since (t,x)V t0f(x) is the solution of (3.8) we get

(3.14) V t0f(Φt(x))=f(Φt(x))+0tΨ(V s0f(Φt(x)))ds.

But by (3.12) we have V t0(fΦt)(x)=V t0f(Φt(x)) and therefore V s0f(Φt(x))=V s0(fΦs)(Φts(x)) , hence Ψ(V s0f(Φt(x)))=Sts(Ψ(V s0(fΦs)))(x) . We conclude that equality (3.14) becomes

V t0(fΦt)(x)=Stf(x)+0tSts(Ψ(V s0(fΦs)))(x)ds,

hence (3.1) is verified as claimed.

To prove (3.7) observe that it is equivalent to

T^tF(μ)=T t0^F(Φt(μ))forallFbp𝒨(E),t0,andμM(E).

Arguing as before, the equality is equivalent to

Vtf=V t0fΦtforallfbp𝒝(E)andt0.

But taking into account (3.12) this last equality holds.

Assertion (ii) is a consequence of Section 4 from [3] , applied to Φ and X0^ on M(E). Notice that as a consequence of (3.12) the kernels T t0^ and Q t0 commute for all t, t′ ⩾ 0.

(iii) The fact that vt given by (1.4) is a solution to equation (3.1) and the second equality of (1.4) follow from (3.2) and (3.7).

Corollary 3.4

Let Ψ be a branching mechanism. Then the following assertions hold.

  1. Consider the pure branching (X0 , Ψ)-superprocess, the right Markov process X0^=(X t0^,𝕇0^μ) with state space M(E), and let u ∈ bp𝒝(E). Let

    V t0u(x):=ln𝔼0^δxeu(X t0^),t0,xE,

    then wt:=V t0u , t ⩾ 0, solves the following nonlinear evolution equation:

    (3.15) ddtwt=Ψ(,wt),t0,

    with the initial condition w0 = u.

  2. Let (𝒧0 , 𝒟(𝒧0)) be the extended weak generator of the pure branching (X0 , Ψ)-superprocess on M(E).

Let u ∈ bp𝒝0, then F := eu belongs to 𝒟(𝒧0), 𝒧0F = −F · lΨ(·,u), and for μM(E) we have

(3.16) 𝒧0F(μ)=Ec(x)F(μ,x)μ(dx)Eb(x)F(μ,x)μ(dx)+E0[F(μ+sδx)F(μ)sF(μ,x)]N(x,ds)μ(dx).

Proof

(i) By (3.2) applied to the transition function of X0^, we know that V t0u, t ⩾ 0, is the solution to the integral equation (3.6). On the other hand, we already observed in the proof of Theorem 3.2 that equation (3.6) is equivalent to (3.8). The assertion follows now by Lemma 3.3 (ii.2); notice that we have K u(x) = Ψ0(x, u(x)), xE, and therefore K preserves the pointwise convergence, the property required by the hypothesis of the lemma.

(ii) Let u𝒝0 , then limt0V t0u(x)=u(x) for all xE. From dominated convergence and (3.2) we get limt0T t0^F(μ)=limt0eμ,V t0u=F(μ) for each μM(E). By Lemma 3.1 (iii) from [3] we conclude that eub𝒝o(𝕋^0).

Let α(t) := 〈μ, wt〉, t ⩾ 0. Arguing as in the proof of Proposition 3.2 from [2], if we fix to > 0 and put λ1 := eβtou we get that ‖wtλ1 for all tto and that Ψ is bounded on E × [0, λ1]. Therefore the map t ↦ 〈μ, Ψ(·, wt)〉 is bounded and continuous on [0, to]. We have wt=f+0tΨ(,ws)ds and thus α(t)=μ,f+0tμ,Ψ(,ws)ds. We deduce that the function α is differentiable on [0, to) and α′(0) = 〈μ, Ψ(·, u)〉. Using again (3.2) we obtain 𝒧0F(μ):=limt0eV t0u(μ)eu(μ)t=ddteα(t)|t=0=eμ,uμ,Ψ(,u), hence 𝒧0F(μ) = −F(μ) · lΨ(·,u)(μ).

We can show now that (3.16) holds. Indeed, the last equality we proved means that 𝒧0F(μ)=F(μ)Eμ(dx)[b(x)u(x)c(x)u(x)2+0(1eu(x)su(x)s)N(x,ds)] and since F′(·, x) = −u(x)F and F″(·, x) = u2(x)F, a straightforward computation leads to (3.16). As in the proof of Proposition 3.1 (ii) we also have T t0^𝒧0F=ewtlΨ(,wt) and letting t ↘ 0, by dominated convergence we get that 𝒧0F𝒝o(𝕋0^). In addition |T^ t0F(μ)F(μ)|tsup(x,λ)E×[0,λ1]|Ψ(x,λ)|l1(μ). Since by (3.3) we have l1𝒠(𝕋0^β), we conclude that F𝒟(𝒧0).

Remark 3.5

(i) According to [22], Theorem (4.1) (a) and its proof, if n ⩾ 1, f1, . . . , fn𝒟(Dw), and ψC 0(n), then the function F on M(E), defined as F(μ) := ψ(〈μ, f1〉, . . . , 〈μ, fn〉), μM(E), belongs to 𝒟(𝒧¯) and the following formula holds

(3.17) 𝒧¯F(μ)=Ec(x)F(μ,x)μ(dx)+E[LF(μ,)(x)b(x)F(μ,x)]μ(dx)+E0[F(μ+sδx)F(μ)sF(μ,x)]N(x,ds)μ(dx).

In the case when the spatial motion is given by a right continuous flow, taking into account (3.5) and (3.16), the equality 𝒧=D^+𝒧0 from assertion (ii) of Theorem 3.2 becomes a version of the above representation (3.17) of 𝒧.

(ii) If E has d points then M(E) is identified with +d and in this case the pure branching (X0, Ψ)-superprocess from Theorem 3.2. with Ψ which is independent of the spatial variable, is precisely a continuous-state branching process (CB-process), as considered and studied in [21, 30].

(iii) Comparing equations (1.3) and (3.15), it follows that a pure branching superprocess has no spatial motion term in the associated evolution equation. Corollary 3.4 (ii) and (3.17) underline the nonlinear PDE counterpart of the pure branching superprocesses.

(iv) The operator D^ occurring in assertion (ii) of Theorem 3.2 is the extended weak generator of a continuous flow on M(E), therefore one can prove that D^(F2)=2FD^F, provided that Fb𝒟 oc(D^) and D^F is a bounded function. Hence D^ acts as a derivation, it is a first order operator in an infinite dimensional frame. Consequently, 𝒧 can be seen as the modification of 𝒧0 with a “drift type” operator and the representation (1.2) may be obtained as a consequence; see [3] for details.

4 Right continuous flows driving non-local branching processes on finite configurations

Let (bk)k⩾1 be a sequence of functions from bp𝒝(E) such that k1bk1, let m1:=k1kbk and assume that 1 < m1 < ∞. We fix also a constant c such that 0<cm1m11.

Recall that the set Ê of all finite configurations of E is identified with the union of all symmetric k-th powers E(k) of E, i.e., if k ⩾ 1 then E(k) is the factorisation of the Cartesian product Ek by the equivalence relation induced by the permutation group σk. Hence E^=k0E(k), where E(0) := {0}; see e.g. [11, 26]. For each k ⩾ 1, let Bk be a Markovian kernel from E(k) to E.

If ϕ ∈ p𝒝(E), 0 ⩽ ϕ ⩽ 1, then by Proposition 4.1 from [8] (see also Theorem 3.1 in [10]), the integral evolution equation

(4.1) ht=ectTtϕ+c0tec(ts)Ttsk1bkBk(h s(k))ds,t0,

has a unique solution tHtϕ, jointly measurable in (t, x) ∈ ℝ+ × E, such that 0 ⩽ Htϕ ⩽ 1. Here, for a function h ∈ bp𝒝(E) we have denoted by h(k), k ⩾ 1, the function on E(k) defined as h(k)(x) := h(x1) · · · h(xk) for all x = (x1, . . . , xk) ∈E(k). We also denote by ĥ: Ê → ℝ+ the function defined as ĥ|E(k) = h(k) if k ⩾ 1 and ĥ(0) = 1.

The integral evolution equation (4.1) associated with a non-local branching process on Ê is formally equivalent to the equation

(4.2) ddtht=(Lc)ht+ck1bkBk(h t(k)),t0,

with the initial condition h0 = ϕ, where L is the generator of the spatial motion; see e.g. Remark 4.2 (ii) from [8].

It turns out that the nonlinear semigroup (Ht)t⩾0 induces a branching semigroup of kernels (Ĥt)t⩾0 on Ê such that H^tϕ^=Htϕ^ for all ϕ ∈ bp𝒝(E), 0 ⩽ ϕ ⩽ 1. According to Theorem 4.1. from [8], under some additional assumptions, there exists a branching right Markov process X^ with state space Ê (depending on the spatial motion X, (bk)k⩾1, the sequence of kernels (Bk)k⩾1, and c) having the transition function (Ĥt)t⩾0.

Non-local pure branching processes on Ê. We consider now the non-local pure branching process X0^ on Ê for which the spatial motion is the trivial Markov process X0. In this case equation (4.1) becomes

(4.3) ht=ectϕ+c0tec(ts)k1bkBk(h s(k))ds,t0.

Let (H t0ϕ)t0 be the solution to equation (4.3), then the transition function of X0^ is (H t0^)t0.

The next proposition emphasises absorbing sets with respect to a pure branching processes on Ê, it is a version of a result from [6].

Proposition 4.1

Let M𝒝(E) and suppose that

(4.4) bk(x)Bk,x(M^E(k))=0foreveryk1andxE\M,

that is, the sub-probability measure on E(k) induced by the sub-Markovian kernel bkBk at x is carried by E(k)\M^. Then E^\M^ is a finely closed absorbing subset of Ê with respect to the pure branching process X0^. The restriction of the pure branching process X0^ to M^ is still a pure branching process, it is induced by the trivial process on M and the restrictions of Bk, k ⩾ 1, to M.

Proof

The assertion follows arguing as in the proof of assertion (iii) of Theorem 3.2 from [6], applied for A := E \ M and observing that A is an absorbing set with respect to the trivial process X0 on E.

For each k ⩾ 1 let (Bk, 𝒟(Bk)) be the operator on bp𝒝(E) defined as Bkϕ := Bk(ϕ(k)), with 𝒟(Bk) := {ϕ ∈ p𝒝(E) : ϕ ⩽ 1}.

Further, as in [7, 25], for the non-local pure branching process X0^=(X t0^)t0 we shall use the notation X t0^=(x t1,,x tN(t))E(N(t)), where N(t) is the number of the particles at time t.

We assume further that each bk, k ⩾ 1, is a constant function on E, so, (bk)k⩾1 is a sequence of positive numbers.

Theorem 4.2

Let X^=(X^t)t0 be the non-local branching process on Ê, depending on a spatial motion given by a right continuous flow Φ = (Φt)t⩾0 on E, (bk)k⩾1, the sequence of kernels (Bk)k⩾1, and c. Assume that the mapping [0, ∞) × E ∋ (t, x) ↦Φt(x) is continuous and suppose that

(4.5) Bk𝒦(Φt)forallt0andk1.

Then the following assertions hold.

  1. The branching process X^=(Xt^,𝕇^μ) has the representation (1.2) on Ê as the pure branching process X0^=(X t0^,𝕇0^μ) driven by the continuous flow Φ and (3.7) holds for all μÊ and Γ𝒝(Ê).

  2. Equation (4.2), with L as the generator of the right continuous flow Φ, has a mild solution, namely, the unique solution ht, t ⩾ 0, to equation (4.1) given by

    (4.6) ht(x)=𝔼0^δxk=1N(t)ϕ(Φt(x tk))=𝔼0^δΦt(x)k=1N(t)ϕ(x tk),xE.

Proof

By (A3.3) from [8], the operator B:=ck1bkBk is Lipschitz; notice that (4.3) is precisely of the form (3.9) from Lemma 3.3 with a := c and condition (II) holds. From (4.5) it follows that we also have B𝒦(Φt) for all t ⩾ 0 and therefore we can apply Lemma 3.3 (iii) to conclude that

(4.7) H t0𝒦(Φt)forallt0.

  1. Arguing as in the proof of Theorem 3.2. to prove (1.2) on Ê we have to show that H^tϕ^(μ)=𝔼0^μϕ^(Φt(X t0^)) for all ϕ ∈ p𝒝(E), ϕ ⩽ 1, and μÊ. Since Ĥt and H t0^ are branching kernels on Ê and ϕ^Φt=ϕΦt^, it is enough to show that

    (4.8) Htϕ=H t0(ϕΦt)forallϕp𝒝(E),ϕ1,andt0.

    By the uniqueness of solutions to equation (4.1), it is sufficient to prove that tH t0(ϕΦt) also satisfies (4.1) with St instead of Tt. Because (H t0ϕ)t0 is the solution to equation (4.3) it follows that

    H t0ϕΦt=ectϕΦt+0tec(ts)B(H s0ϕ)Φtdsforallt0.

    Since H t0 and B belong to 𝒦(Φt′), t′ ⩾ 0, and using property (p2) from Lemma 3.3, the last equality becomes

    H t0(ϕΦt)=ectStϕ+0tec(ts)StsB(H s0(ϕΦs))dsforallt0,

    which means that indeed (H t0(ϕΦt))t0 is a solution to (4.1) and therefore (4.8) holds and the proof of assertion (i) is complete.

  2. The first equality of (4.6) follows from (1.2) since (Ĥ t)t⩾0 is the transition function of X^, while the second one is a consequence of (4.8) and (4.7).

Remark 4.3

  1. Example of kernels satisfying condition (4.5). For each k ⩾ 1 define the kernel Bk as BkF(x) := F(x, . . . , x) for all xE and F ∈ bp𝒝(E(k)). Then B kϕ = ϕk for all ϕ ∈ p𝒝(E), ϕ ⩽ 1, and clearly (4.5) is verified. Notice that equation (4.1) becomes

    ht=ectTtϕ+c0tec(ts)Tts(k1bkh sk)ds,t0.

  2. Recall the classical probabilistic meaning of the branching procedure: if k ⩾ 1 and xE, then bk(x) represents the probability that a particle destroyed at the point xE has precisely k descendants, and Bk,x is the distribution of the k descendants. As in [6], taking into account this probabilistic description of a branching process, one can explain the additional condition from Proposition 4.1, imposed to the kernels Bk: (4.4) means that when a particle is destroyed at a point of E \ M then not all the direct descendants are distributed inside M.

  3. Consider the classical kernels Bk, k ⩾ 1, from assertion (i) and observe that condition (4.4) is satisfied in this case for every set M𝒝(E). Consequently, applying Proposition 4.1 we get that: the set E^\M^ is absorbing with respect to any pure branching process X0^ on Ê, for all M𝒝(E). In particular, we highlighted absorbing sets for the induced non-local pure branching process from [7] , occurring in the study of the nonlinear evolution equation considered there.

  4. We can state for non-local branching processes on Ê results similar to (1.4) and Corollary 3.4. Let X0^=(X t0^,𝕇0^μ) be the non-local pure branching process on Ê, for which the spatial motion is the trivial Markov process X0, and let ϕ ∈ p𝒝(E), ϕ ⩽ 1. If we put

    ht(x):=𝔼0^δxϕ^(X t0^),xE,t0,

    then ht, t ⩾ 0, solves the nonlinear evolution equation

    (4.9) ddtht=cht+ck1bkBk(h t(k)),t0,

    with the initial condition h0 = ϕ; observe that equation (4.9) is precisely (4.2) with L = 0. Indeed, ht, t ⩾ 0, is the solution to the integral equation (4.3). The assertion follows by Lemma 3.3 (ii.2), applied for Kϕ=ck1bkBk(ϕ(k)) and observing that K preserves the pointwise convergence.

  5. In [28] it is studied an equation analogue to (4.9) (see also [27]) on the locally finite configurations of a locally compact metrizable space, however, with no associated branching Markov process.

5 Nonlinear evolution equations on measures

Let (bk)k⩾1 be a sequence of positive numbers such that k1bk1 and 1<k1kbk< as in the previous section. Let Y0^=(Y t0^)t0 be the non-local pure branching process from Remark 4.3 (i) with the state space M(E)^ instead of Ê, so, it depends on the sequence (bk)k⩾1, the constant c, and its spatial motion is the trivial Markov process Y0 on M(E). The corresponding equation (4.1) becomes

ht=ectf+c0tec(ts)k1bkh skdsonM(E),t0.

Let (L, 𝒟(L)) be the extended weak generator of a spatial motion X on E and let 𝕋 = (Tt)t⩾0 be its transition function.

We consider on M(E) the following nonlinear evolution equation

(5.1) ddtFt=LFtcFt+ck1bkF tkonM(E),t0,

with the initial condition F0 = F, where for a function F : M(E) → [0, 1] we denoted by F′ its variational derivative and LF(μ):=ELF(μ,x)μ(dx).

Let Φ^=(Φ^t)t0 be the right continuous flow on M(E) induced by the right Markov process X, Φ^t(μ):=μTt=Xt(𝕇μ), μM(E). Recall that by Proposition 3.1 (ii) the extended weak generator of Φ^ coincides with FLF′ on a large class of functions F on M(E).

Further, we need to consider the following condition:

(5.2) themapping[0,)×M(E)(t,μ)μTtM(E)iscontinuous

Let 𝒢 ⊂ bC(E) be a family of functions which is convergence determining, i.e., for every sequence (μn)n of probabilities on E we have that (μn)n is weakly convergent to μ if and only if μn(g) → μ(g) for all g𝒢; see [20] , page 112.

Remark 5.1

The following Feller type property of the transition function 𝕋 implies that condition (5.2) holds: There exists a convergence determining family 𝒢 such that Tt(𝒢) ⊂ bC(E) for every t > 0 and limt↘0Ttgg = 0 for all g𝒢. In particular, if E is a locally compact space with countable base and 𝕋 = (Tt)t⩾0 is a Feller semigroup on E, that is, Tt maps C0(E) into itself for every t > 0, then (5.2) is satisfied.

The next corollary follows from Theorem 4.2 (ii) and the above considerations.

Corollary 5.2

Equation (1.1) has a mild solution Ft, t ⩾ 0, with F0 = F ∈ p𝒨(ℝd), F ⩽ 1, uniquely determined by the pure branching process Y0^ on M(d)^, the d-dimensional Brownian motion (Bt)t⩾0, and the constant c,

(5.3) Ft(μ)=𝔼0^δμk=1N(t)F(Bt(𝕇y tk))=𝔼0^δBt(𝕇μ)k=1N(t)F(y tk),μM(d),

where Y t0^=(y t1,,y tN(t)) and N(t) is the number of the particles in M(ℝd) at time t.

More generally, if we assume that condition (5.2) holds then equation (5.1) has a mild solution uniquely determined by the pure branching process Y0^ on M(E)^, the spatial motion X, and the constant c, namely, for every F ∈ p𝒨(E), F ⩽ 1, the nonlinear evolution equation on measures

Ft=ectFΦ^t+c0tec(ts)k1bkF skdsonM(E),t0,

has a unique solution tFt, jointly measurable in (t, μ) ∈ ℝ+ × M(E), such that 0 ⩽ Ft ⩽ 1 and (5.3) holds with Xt instead of Bt and M(E) instead of M(ℝd).

Remark

In [12] , in solving a nonlinear Schrödinger equation of a non-local branching process, the solution is obtained from an Lp -weak solution for a stationary linear equation similar to (5.1), associated to a branching semigroup on Ê and using a Feynman-Kac formula stated in [5] .

Appendix

(A1) Proof of Proposition 3.1. (i) Let ub𝒟 oc(L) be such that Lu is a bounded function and tLTtu(x) is continuous. Replacing 𝕋b by 𝕋b+α, we may assume that u = Ubf with f bounded and tT tbf(x) continuous, where Ub is the potential kernel of 𝕋b, Ub:=0T tbdt. Then we have T tbuut=1t0tT sbfds and |T tbuut|f By dominated convergence we obtain limt0ET tbuutdμ=μ(f) for every μM(E). From (3.4) we have |T^t(lu)lut|=|lT tbuut|fl1. On the other hand from (3.3) we get Ûβl1 < ∞ for some β > 0. We deduce that lu𝒟(𝒧) and 𝒧(lu) = lLubu.

(ii) Observe first that if u ∈ b𝒟(L) is such that Lu is bounded and we fix α > 0 then u = Uαg for some function g ∈ b𝒝0(𝕋). If μM(E) then by dominated convergence limt0μ,Ttuut=lLu(μ). Let u := (u1, . . . , un), Ttu := (Ttu1, . . . , Ttun), Lu := (Lu1, . . . , Lun), and lu := (lu1, . . . , lun). Hence limt0Q t0F(μ)=limt0ψ(lTtu(μ))=ψ(lu(μ))=F(μ), F ∈ b𝒝0(𝕈0) = b𝒝o(𝕈0). We have also limt0Q t0F(μ)F(μ)t=ψ(lu(μ))lLu(μ)=μ,LF(μ,)=:G(μ). Then Q t0G(μ)=ψ(lTtu(μ))lLTtu(μ) and therefore G𝒝o(𝕈0). It remains to show that F𝒟(D^). Let to > 0 such that for all t < to we have TtuutC1 for some constant C1 > 0. Since ψ is Lipschitz we get |Q t0F(μ)F(μ)t|C2|lTtu(μ)lu(μ)|tC2Ttuutμ(1)C2C1l1(μ) for all t < to and μM(E), with C2 > 0. By (3.3), arguing as in last part of the proof of assertion (i) we conclude that F𝒟(D^) and D^F(μ)=μ,LF(μ,).

(A2) Proof of Lemma 3.3. The proof of assertion (i) is straightforward and we omit it.

(ii.1) We argue as in the proof of Proposition 3.2 from [2] and show the existence of a solution to equation (3.9) by Picard approximation. Let w t0=0 and for k ⩾ 0 and xE define inductively

w tk+1(x)=eatf(x)+0tea(ts)K(w sk)(x)ds,0tto.

Observe first that if condition (II) holds and we take f ∈ bp𝒝(E), f ⩽ 1, then we have w tk1 for all k ⩾ 0. Indeed, we reason by induction. If we suppose that w sk1 then, since Kfa, we get w tk+1eat+a0teasds=1. So, in this case w sk belongs to 𝒟(K) for all s and k.

One can see that 0w tk(x)eCtof for all xE and k ⩾ 0, where C is the Lipschitz constant of K. We have also w tk+1w tk(Ct)kk!f and therefore the limit limkw tk(x)=:Wtf(x) exists uniformly in (t, x) ∈ [0, to] × E and in particular, tWtf is a solution to (3.9).

The uniqueness of the solution to (3.9) follows by Gronwall’s lemma, using the Lipschitz property of K . (ii.2) First, using the hypothesis that K preserves the pointwise convergence, one can prove inductively that for every k ⩾ 0 and xE the function [0,to]tw tk(x) is continuous. Since ut=limkw tk uniformly in t, it follows that tut(x) is also continuous on [0, to]. Consequently, (3.10) is now a direct consequence of (3.9).

(iii) By the properties (p3) and (p2) of 𝒦(J) and since we assumed that K𝒦(J), one can see inductively that w tk𝒦(J) for all k and t. We have Wtf(x)=limkw tk(x) for all xE and therefore, by property (p1) of 𝒦(J) we deduce that Wt𝒦(J) for all t.

(A3) Right continuous flow. Assume that E is a Lusin topological space and 𝒝(E) is the Borel σ-algebra. A family Φ = (Φt)t⩾0 of mappings on E is called right continuous flow (cf. [32] , page 41) provided that:

  1. Φt+s(x) = Φt(Φs(x)) for all s, t > 0 and xE;

  2. Φ0(x) = x for all xE;

  3. For each t > 0 the function ExΦt(x) is 𝒝(E)/𝒝(E)-measurable;

  4. For each xE the function tΦt(x) is right continuous on [0, ∞).

(A4) The extended weak generator. We recall several facts about the weak generator of a right Markov process X with state space E, transition function 𝕋 = (Tt)t⩾0, and 𝒰 = (Uα)α>0 as associated resolvent of kernels; for details see [3], where the classical weak generator acting on bounded functions, considered by E.B. Dynkin in [17], was extended to unbounded real-valued functions.

Define 𝒝0 = 𝒝0(𝕋):= {f ∈ [𝒝(E)] : Tt(|f|) < ∞ for all t > 0 and f =lims↘0 Tsf pointwise on E}. Clearly, we have [𝒠(𝒰α)] ⊂ 𝒝0 for every α ⩾ 0 and every bounded finely continuous function belongs to 𝒝0, in particular, bC(E) ⊂ 𝒝0.

Define also 𝒝o = 𝒝o(𝕋) := {f𝒝0 : for each α > 0 there exist to > 0 and hα ∈ p𝒝(E) such that sup0<s<to Ts|f|hα with Uαhα < ∞}.

Let 𝒟(L) be the set of all u𝒝o such that for each α > 0 there exist to > 0 and hα ∈ p𝒝(E) with sup0<t<to|Ttuut|hα, Uα hα < ∞ on E, and there exists pointwise limt0Ttuut𝒝o. Clearly, 𝒝o and 𝒟(L) are vector spaces and define the linear operator L : 𝒟(L) → 𝒝o as Lu(x):=limt0Ttu(x)u(x)t, f𝒟(L), xE.

The operator (L, 𝒟(L)) is called the extended weak generator of 𝕋 = (Tt)t⩾0 and one can prove that if α > 0 then 𝒟(L) = Uα(𝒝o).

In [24] it was considered an extended generator (L¯,𝒟(L¯)) of 𝕋 = (Tt)t⩾0, taking into account unbounded real-valued functions, as follows: let u, g𝒝0, then u belongs to the domain 𝒟(L¯) of L¯ and g=L¯u provided that for all t > 0 and xE we have 0tTs(|g|)(x)ds< and Ttu(x)=u(x)+0tTsg(x)ds. According to [24], Proposition 4.1 (see also [22], page 354, the proof of Theorem (4.1)), we have the following equivalent definition for the extended generator: if u, g𝒝0 then u𝒟(L¯) and L¯u=g if and only if for all xE we have stTs(|g|)(x)ds< for all t > 0 and (u(Xt)u(X0)0tg(Xs)ds)t0 is a (𝒡t)-martingale under 𝕇x.

Let 𝒝oo := {f𝒝o: for each t > 0 there exist to > 0 and ht ∈ p𝒝 such that sup0<s<to Ts|f|ht with Ttht < ∞} and define 𝒟o(L) := {u𝒟(L) : Lu𝒝oo} and 𝒟 oc(L):={u𝒟o(L):[0,)tLTtu(x) is continuous for each xE}.

Acknowledgment

This work was supported by a grant of Ministry of Research and Innovation, CNCS - UE-FISCDI, project number PN-III-P4-PCE-2021-0921, within PNCDI III. The second named author kindly acknowledges support from eMAG.

  1. Conflict of interest statement:

    Authors state no conflict of interest.

References

[1] V. Barbu and L. Beznea, Measure-valued branching processes associated with Neumann nonlinear semiflows, J. Math. Anal. Appl.441 (2016), 167–182.10.1016/j.jmaa.2016.04.005Search in Google Scholar

[2] L. Beznea, Potential theoretical methods in the construction of measure-valued branching processes, J. Eur. Math. Soc.13 (2011), 685–707.10.4171/JEMS/264Search in Google Scholar

[3] L. Beznea, M. Bezzarga, and I. Cîmpean, Markovian flows driving Markov processes and multiplicative Lp-semigroups, preprint (2022).Search in Google Scholar

[4] L. Beznea and N. Boboc, Potential Theory and Right Processes. (Mathematics and Its Applications, vol. 572), Kluwer Academic Publishers/Springer, 2004.10.1007/978-1-4020-2497-9Search in Google Scholar

[5] L. Beznea and N. Boboc: Feynman-Kac formula for left continuous additive functionals and extended Kato class measures. Potential Anal.30 (2009), 139–164.10.1007/s11118-008-9109-1Search in Google Scholar

[6] L. Beznea, A.-M. Boeangiu, and O. Lupaşcu-Stamate, h-transform of Doob and nonlocal branching processes, Anal. Math. Phys. (2020) 10:47, https://doi.org/10.1007/s13324-020-00390-310.1007/s13324-020-00390-3Search in Google Scholar

[7] L. Beznea, L.I. Ignat, and J. D. Rossi: From Gaussian estimates for nonlinear evolution equations to the long time behavior of branching processes. Rev. Mat. Iberoam.35 (2019), 823–84610.4171/rmi/1071Search in Google Scholar

[8] L. Beznea and O. Lupaşcu, Measure-valued discrete branching Markov processes, Trans. Amer. Math. Soc.368 (2016), 5153–5176.10.1090/tran/6514Search in Google Scholar

[9] L. Beznea, O. Lupaşcu, and A.-G. Oprina, A unifying construction for measure-valued continuous and discrete branching processes, in: Complex Analysis and Potential Theory, CRM Proceedings and Lecture Notes, vol. 55, Amer. Math. Soc., Providence, RI, 2012, 47–59.10.1090/crmp/055/04Search in Google Scholar

[10] L. Beznea, O. Lupaşcu-Stamate, and C.I. Vrabie, Stochastic solutions to evolution equations of non-local branching processes, Nonlinear Anal.200 (2020), 112021. https://doi.org/10.1016/j.na.2020.11202110.1016/j.na.2020.112021Search in Google Scholar

[11] L. Beznea and A.-G. Oprina, Nonlinear PDEs and measure-valued branching type processes, J. Math. Anal. Appl.384 (2011), 16–32.10.1016/j.jmaa.2010.10.034Search in Google Scholar

[12] L. Beznea and A.-G. Oprina, Bounded and Lp-weak solutions for nonlinear equations of measure-valued branching processes, Nonlinear Anal.107 (2014), 34–4610.1016/j.na.2014.04.020Search in Google Scholar

[13] L. Beznea and M. Röckner, From resolvents to càdlàg processes through compact excessive functions and applications to singular SDE on Hilbert spaces, Bull. Sci. Math.135 (2011), 844–870.10.1016/j.bulsci.2011.07.002Search in Google Scholar

[14] D. J. Daley and D. Vere-Jones, An Introduction to the Theory of Point Processes. Vol. I. Elementary Theory and Methods (second edition), Springer, 2003Search in Google Scholar

[15] D.A. Dawson, Measure-valued Markov processes, in: École d’Été de Probabilités de Saint-Flour XXI–1991 (Lecture Notes in Math. 1541), Springer (1993), 1–260.10.1007/BFb0084190Search in Google Scholar

[16] D.A. Dawson, L.G. Gorostiza, and Z.H. Li, Nonlocal branching superprocesses and some related models, Acta Appl. Math.74 (2002), 93–112.10.1023/A:1020507922973Search in Google Scholar

[17] E. B. Dynkin, Markov Processes, Vol. I, Springer, 1965.10.1007/978-3-662-00031-1Search in Google Scholar

[18] E. B. Dynkin, Diffusions, superdiffusions and partial differential equations, Amer. Math. Soc., Colloquium Publications, Vol. 50, 2002.10.1090/coll/050Search in Google Scholar

[19] N. El Karoui and S. Roelly, Propriétés de martingales, explosion et représentation de Lévy-Khintchine d’une classe de processus de branchement à valeurs mesures, Stochastic Process. Appl.38 (1991), 239–266.10.1016/0304-4149(91)90093-RSearch in Google Scholar

[20] S.N. Ethier and T.G. Kurtz, Markov Processes: Characterization and Convergence, Wiley & Sons, 1986.10.1002/9780470316658Search in Google Scholar

[21] R. Fang and Z. Li, A conditioned continuous-state branching process with applications, Statist. Probab. Lett.152 (2019), 43–49.10.1016/j.spl.2019.04.013Search in Google Scholar

[22] P. J. Fitzsimmons, Construction and regularity of measure-valued Markov branching processes, Israel J. Math.64 (1988), 337–361.10.1007/BF02882426Search in Google Scholar

[23] T. E. Harris, The Theory of Branching Processes, Springer, 1963.10.1007/978-3-642-51866-9Search in Google Scholar

[24] F. Hirsch and M. Yor, On temporally completely monotone functions for Markov processes, Probab. Surv.9 (2012), 253–286.10.1214/11-PS179Search in Google Scholar

[25] P. Hsu, Branching Brownian motion and the Dirichlet problem of a nonlinear equation, in: Seminar on stochastic processes, 1986, Progr. Probab. Statist. 13, Birkhäuser (1987), 71–83.10.1007/978-1-4684-6751-2_6Search in Google Scholar

[26] N. Ikeda, M. Nagasawa, and S. Watanabe, Branching Markov processes. I J. Math. Kyoto Univ.8 (1968), 233–278.10.1215/kjm/1250524059Search in Google Scholar

[27] Y. Kondratiev and Y. Kozitsky, The evolution of states in a spatial population model, J. Dynam. Differential Equations30 (2018), 135–173.10.1007/s10884-016-9526-6Search in Google Scholar

[28] Y. Kozitsky and A. Tanaś, Evolution of states of an infinite particle system with nonlocal branching, J. Evol. Equ. (2022), to appear.10.1007/s00028-022-00775-wSearch in Google Scholar

[29] J.-F. Le Gall, Spatial Branching Processes, Random Snakes and Partial Differential Equations (Lectures in Mathematics ETH Zürich), Birkhäuser, 1999.10.1007/978-3-0348-8683-3Search in Google Scholar

[30] Z. Li, Measure-Valued Branching Markov Processes, Probability and its Applications (New York), Springer, 2011.10.1007/978-3-642-15004-3Search in Google Scholar

[31] M. Nagasawa, Schrödinger Equations and Diffusion Theory, Birkhäuser, 1993.10.1007/978-3-0348-0560-5Search in Google Scholar

[32] M. Sharpe, General Theory of Markov Processes, Academic Press, Boston, 1988.Search in Google Scholar

Received: 2021-07-09
Accepted: 2022-01-04
Published Online: 2022-02-25

© 2022 Lucian Beznea et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 12.6.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2021-0229/html
Scroll to top button