arXiv:1104.1496v1 [math.PR] 8 Apr 2011 o rcs,cniinn,rno environments. random conditioning, process, Cox representati diffusion, measure-valued process, aino h iiigifiiepril ytm oto thes of Most system. particle [ infinite proce limiting measure-valued the the of of mation parti representation a distinct give preserve which which systems. and particle developed finite been of have measures proaches empirical of limits tem tt ftemauevle rcs samlil fted Fi de the of multiple a as process measure-valued the of state 939–984 3, No. DOI: 39, Vol. Probability2011, of Annals The

c 5 n yorpi detail. 939–984 typographic 3, the and by No. published 39, Vol. article 2011, original the Mathematical of of reprint Institute electronic an is This OSO ERSNAIN FBACIGMRO AND MARKOV BRANCHING OF REPRESENTATIONS POISSON – nttt fMteaia Statistics Mathematical of Institute .Introduction. 1. e od n phrases. and words Key 2 1 M 00sbetclassifications. subject 2000 AMS 2009. November revised 2008; January Received 7 upre npr yDAAUA,Mxc,Gat968SFA. Grant Mexico, DMS-08-057 DGAPA-UNAM, and by DMS-05-03983 part Grants in NSF Supported by part in Supported 10.1214/10-AOP574 , 35 aeepotdpoete fecagal eune,ide sequences, exchangeable of properties exploited have ] n rcse,addffso prxmtosfrpoessi processes for approximations environments. diffusion and processes, supercri ing for e theorem on convergence conditioning Harris’s including nonextinction, results and calculations of ety process. measure-valued desired the is iiigmauevle rcs,a ahtime each at process, measure-valued limiting ino oain n eesi odtoal oso distr Poisson conditionally is measure levels mean and locations of tion rfrtelmtn oes isifiiy o rnhn Mark branching For infinity. each hits at models, cesses, limiting the for or eesaeidpnetaduiomydsrbtdo [0 on distributed uniformly and independent are levels cusol hntelvlo h atcecossaspecified p a a crosses particle of the earlie death of unlike fact, level In but the r time. “level,” when death with only a change occurs and are and levels birth location the varying a of structions, has spatially systems particle with countable Each models of for terms structed in limits valued yToa .Kurtz G. Thomas By h ersnainsmlfiso ie lentv rosf proofs alternative gives or simplifies representation The ESR-AUDBACIGPROCESSES BRANCHING MEASURE-VALUED ersnain fbacigMro rcse n hi me their and processes Markov branching of Representations nvriyo icni,MdsnadUNAM and Madison Wisconsin, of University esr-audpoessaientrlya nnt sys- infinite as naturally arise processes Measure-valued K rnhn akvpoes asnWtnb rcs,super process, Dawson–Watanabe process, Markov Branching ( t ) × 2011 , t odtoe ntesaeo h rcs,the process, the of state the on conditioned , hsrpitdffr rmteoiia npagination in original the from differs reprint This . Λ where , 02,6K5 06,6J0 60K37. 60J80, 60J60, 60K35, 60J25, 1 in n laeR Rodrigues R. Eliane and Λ h naso of Annals The 1 eoe eegemaue and measure, Lebesgue denotes n elrdffso,exchangeability, diffusion, Feller on, t h on distribu- joint the , ia branch- tical r , tnto or xtinction btdwith ibuted .Frthe For ]. random n representations e ubro ap- of number A ravari- a or lsi h limit the in cles sa transfor- a as ss , 93. vpro- ov level article con- r et measure netti asure- ates. con- tfigthe ntifying 2 K r , - 2 T. G. KURTZ AND E. R. RODRIGUES of the sequence, or, as in [35], as a transformation of the de Finetti measure of a sequence that gives both a location and a mass for the distinct particles. The primary limitation of the representations given for measure-valued branching processes in [7] is that the branching rates must be independent of particle location. This restriction was relaxed in [31] for critical and subcriti- cal branching, but that representation does not seem to provide much useful insight and would be difficult to extend to supercritical processes. A second limitation of the approach in [7] is that, at least without major additional effort, it applies only to models in which the states are finite measures. In the present paper, we give another construction which, although similar to that of [31], applies immediately to both subcritical and supercritical processes (as well as processes that are subcritical in some locations and supercritical in others). The construction also applies immediately to models with infinite mass. In addition, the new construction seems to be a much more effective tool for analyzing the measure-valued processes obtained. We introduce the basic ideas of the construction in Section 2, giving re- sults for population models without location or type. As in the earlier work, the justification for the representation is a consequence of a Markov map- ping theorem, Theorem A.15. The Feller diffusion approximation for nearly critical branching processes is obtained as a consequence of the construction. Section 3 gives the construction for the general branching Markov process and the Dawson–Watanabe limit. Section 4 gives a variety of applications and extensions, including conditioning on nonextinction, mod- els with heavy-tailed offspring distributions and processes in random envi- ronments. The Appendix contains background material and a number of technical lemmas.

2. Simple examples. In this section, we give particle representations for pure death and continuous time Markov branching processes and illustrate how the infinite system limit can be derived immediately. The main is to introduce the notion of the level of a particle in the simplest possible settings.

2.1. Pure death processes. For r> 0, let ξ1(0),...,ξn0 (0) be independent random variables, uniformly distributed on [0, r]. For b> 0, let

ξ˙i(t)= bξi(t), so ξ (t)= ξ (0)ebt, and define N(t)=# i : ξ (t) < r and U(t) = (U (t),..., i i { i } 1 UN(t)), where the Uj(t) are the values of the ξi(t) that are less than r. The Ui(t) will be referred to as the levels of the particles. The level of a particle being below r means that the particle is “alive,” and as soon as its level reaches r the particle “dies.” Note that N(t) is the number of particles “alive” in the system at time t. POISSON REPRESENTATIONS 3

Let f(u, n)= n g(u ), where 0 g 1, g is continuously differentiable i=1 i ≤ ≤ and g(ui)=1 for ui > r. [The “n” in f(u, n) is, of course, redundant, but it will help clarifyQ some of the later calculations.] Then d f(U(t),N(t)) = Af(U(t),N(t)), dt where n Af(u, n)= f(u, n) buig′(ui)/g(ui). Xi=1 Note that Af(u, n) may also be written as n Af(u, n)= buig′(ui) g(uj ). i=1 j=i X Y6 Hence, even if g(ui)=0 for some i, the expression for Af(u, n) still makes sense. Let αr(n, du) be the joint distribution of n i.i.d. uniform [0, r] random λg 1 r λgn variables. Setting e− = r− 0 g(z) dz, f(n)= f(u, n)αr(n, du)= e− and R R b r λg (n 1) 1 Af(u, n)αr(n, du)= ne− − br− zg′(z) dz Z Z0 r λg (n 1) 1 = bne− − r− (g(r) g(z)) dz − Z0 λg (n 1) λg = bne− − (1 e− ) − = Cf(n), where b Cf(n)= bn[f(n 1) f(n)], − − that is, the generator of a linear death process. Of course, the conditional b b b distribution of U(t) given N(t) is just αr(N(t), ). Let = σ(U(s) : s t) and N = σ(N(s) : s · t). Then trivially, Ft ≤ Ft ≤ t f(U(0),N(0)) = f(U(t),N(t)) Af(U(s),N(s)) ds − Z0 is an -martingale, and Lemma A.13 implies {Ft} t E[f(U(t),N(t)) N ] E[Af(U(s),N(s)) N ] ds |Ft − |Fs Z0 t = f(N(t)) Cf(N(s)) ds − Z0 b b 4 T. G. KURTZ AND E. R. RODRIGUES

N is a t -martingale. Consequently, N is a solution of the martingale prob- lem{F for C} and hence is a linear death process. Of course, this observation r r bτi follows immediately from the fact that τi defined by Ui(0)e = r is expo- nentially distributed with parameter b, but the martingale argument illus- trates a procedure that works much more generally.

2.2. A simple . For f and g as above, a> 0, r> 0 and < b ra, define the generator −∞ ≤ n r A f(u, n)= f(u, n) 2a (g(v) 1) dv r − i=1 Zui (2.1) X n 2 g′(ui) + f(u, n) (aui bui) . − g(ui) Xi=1 We refer to ui as the level of the ith particle, and as in the pure death ex- ample, a particle “dies” when its level reaches r. The process with generator (2.1) has the following properties. The particle levels satisfy U˙ (t)= aU 2(t) bU (t), i i − i and a particle with level z gives birth at rate 2a(r z) to a particle whose initial level is uniformly distributed between z and− r. Uniqueness for the martingale problems for Ar follows by first checking uniqueness for the op- erator D given by the second term alone, and then showing that uniqueness holds up to the first time that the process includes n particles by observing that Ar truncated at n is a bounded perturbation of D. Finally, the first hitting time of n goes to infinity as n (cf. Problem 28 in Section 4.11 of [9]). →∞ As before, f(n)= f(u, n)α (n, du)= e λg n. To calculate A (fu, n) r − r × αr(n, du), observe that R R r b r r 1 2λg 1 r− 2a g(z) (g(v) 1) dv dz = are− 2ar− g(z)(r z) dz − − − Z0 Zz Z0 and r r 1 2 1 r− (az bz)g′(z) dz = r− (2az b)(g(z) 1) dz − − − − Z0 Z0 r 1 λg = 2ar− zg(z) dz + ar + b(e− 1). − − Z0 Then

λg (n 1) 2λg λg λg Af(u, n)α (n, du)= ne− − (are− 2are− + ar + b(e− 1)) r − − Z = Cf(n),

b POISSON REPRESENTATIONS 5 where (2.2) Cf(n)= ran(f(n + 1) f(n)) + (ra b)n(f(n 1) f(n)) − − − − is the generator of a branching process. b b b b b Unlike the linear death example, it is not immediately obvious that N αr(N(t), ) is the conditional distribution of U(t) given t = σ(N(s) : s t); however,· Theorem A.15 and the fact that a solutionF of the martingale≤ problem for C starting from N(0) exists gives the existence of a solu- tion (U(t),N(t)) of the martingale problem for Ar such that for all t 0, N ≥ αr(N(t), ) is the conditional distribution of U(t) given t . To apply Theo- rem A.15·, take ψ(u, n)= n in (A.15) and assume E[N(0)]F< . Any solution of the martingale problem for C with E[N(0)] < will satisfy∞ E[N(t)] = E[N(0)]ebt, for all t 0. The moment assumption∞ can be eliminated by conditioning. ≥ We conclude that for any distribution for N(0), there is a solution (U, N) of the martingale problem for A such that N is a solution of the martin- gale problem for C, that is, N is a linear birth and death process with birth rate ar and death rate ar b. Uniqueness holds for the martingale problem for A, so for any solution− of the martingale problem for A satis- fying P U(0) Γ N(0) = αr(N(0), Γ), we have that N is a solution of the martingale{ problem∈ | for}C. This representation can be used to do simple calculations. For example, let U (0) be the minimum of U1(0),...,UN(0). Then for all t, all levels are above∗ bt U (0)e− ∗ U (t)= bt . ∗ 1 (a/b)U (0)(1 e− ) − ∗ − Let τ = inf t : N(t) = 0 . Then if τ is finite, U (τ)= r. In particular, if N(0) = n, then{ } ∗ r P τ>t = P U (t) < r = P U (0) < { } { ∗ } ∗ e bt (ra/b)(e bt 1)  − − − −  n bt ra bt − = 1 e− (e− 1) . − − b −   bt ra bt Note that the assumption that b ra ensures that e− b (e− 1) 1. ≤ − − ≥ 1 In the branching process, the lifetime of an individual is (ar b)− which will be small if r is large. Consequently, it is important to note− that the levels do not represent single individuals in the branching process but whole lines of descent. For example, at least in the critical or subcritical case, the individual with level U (0) at time zero is the individual whose of descent lasts longer than that∗ of any other individual alive at time zero. 6 T. G. KURTZ AND E. R. RODRIGUES

2.2.1. Conditioning on nonextinction. If b 0, then conditioning on non- extinction, that is, conditioning on τ >t and≤ letting t , is equivalent →∞ to conditioning on U (0) = 0. Conditioned on U (0) = 0, U1(0),...,UN(0)(0) include N(0) 1 independent,∗ uniform [0, r] random∗ variables and one that equals zero. If− one of the initial levels is zero, then the solution of the mar- tingale problems for Ar gives a solution for

n 1 r n 1 c − − 2 g′(ui) Arf(u, n)= f(u, n) 2a (g(v) 1) dv + f(u, n) (aui bui) ui − − g(ui) Xi=1 Z Xi=1 r + f(u, n)2a (g(v) 1) dv, − Z0 c and taking αr(n, du) to be the distribution of n independent random vari- ables, one of which is zero and the others uniform [0, r], we see that N is a solution of the martingale problem for Ccf(n)= ra(n + 1)(f(n + 1) f(n)) + (ra b)(n 1)(f(n 1) f(n)). − − − − − 2.2.2.b Conditioningb on extinction.b If 0 < b < ra, then conditioningb b on ex- b tinction is equivalent to conditioning on U (0) > a . Conditioned on U (0) > b ∗ b ∗ a , U1(0),...,UN(0)(0) are independent uniform [ a , r]. Defining Vi(t)= Ui(t) b − a , V is a solution of the martingale problem for n r b/a n − 2 g′(vi) Arf(v,n)= f(v,n) 2a (g(z) 1) dz + f(v,n) (avi + bvi) , vi − g(vi) Xi=1 Z Xi=1 so N is a solution of the martingale problem for Cf(n) = (ra b)n(f(n + 1) f(n)) + ran(f(n 1) f(n)), − − − − which is the generator of a subcritical branching process. b b b b b 2.2.3. Convergence as t . Again, in the supercritical case, 0

If N(0) = 1, then U (0) is uniformly distributed on [0 , r], and hence P W > 0 = b and ∗ { } ra λW 1 E[e− W > 0] = , | 1 + (ra/b)λ b that is, W is exponentially distributed with parameter ra . Of course, we have simply rederived a classical result of Harris [17].

2.3. Feller diffusion approximation. As r , Arf in (2.1) converges for every continuously differentiable g such that→∞ 0 g 1 and g(z)=1 for z r , that is, for f(u)= g(u ), in the limit ≤ ≤ ≥ g i i rg Q 2 g′(ui) (2.4) Af(u)= f(u) 2a (g(v) 1) dv + f(u) (aui bui) . ui − − g(ui) Xi Z Xi If n/r y as r , then α (n, ) converges to α(y, ), where α(y, ) is the → →∞ r · · · distribution of a Poisson process ξy on [0, ) with intensity y, in the sense that ∞

∞ ∞ R log g(z) dξy(z) y R (1 g(z)) dz f(u)αr(n, du) E[e 0 ]= e− 0 − . n → Z[0,r] 8 T. G. KURTZ AND E. R. RODRIGUES

Note that ∞ y R (1 g(z)) dz yβg f(y)= αf(y)= f(u)α(y, du)= e− 0 − = e− , Z and usingb Lemma A.3

yβg ∞ ∞ αAf(y)= e− 2ay g(z) (g(v) 1) dv dz −  Z0 Zz ∞ 2 + y (az bz)g′(z) dz − Z0 

yβg ∞ ∞ = e− 2ay g(z) (g(v) 1) dv dz −  Z0 Zz y ∞(2az b)(g(z) 1) dz − − − Z0 

yβg ∞ ∞ = e− 2ay g(z) (g(v) 1) dv dz −  Z0 Zz 2ay ∞ ∞(g(v) 1) dv dz − − Z0 Zz + by ∞(g(z) 1) dz − Z0  yβg 2 = e− (ayβ byβ ) g − g = Cf(y), where b Cf(y)= ayf ′′(y)+ byf ′(y). Again, we can apply Theorem A.15 taking b b b 1 1 γ(u) =limsup [0,z](ui) z z →∞ Xi ui and ψ(u)= i e− , so 2 rg Af(u) [2ar + g′ (ar + b r )]e ψ(u) P | |≤ g k k g | | g and ψ(u)α(y, du)= y. If Y is a solution of the martingale problem for C bt withRE[Y (0)] < , then E[Y (t)] = e E[Y (0)] and the conditions of Theorem A.15 are satisfied.∞ Consequently,e there is a solution U of the martingale probleme for A such that e e 1 1 (2.5) Y (t) =limsup [0,z](Ui(t)) z z →∞ Xi POISSON REPRESENTATIONS 9 is a solution of the martingale problem for C with the same distribution as Y . [Note that, with probability one, the lim sup in (2.5) is actually a limit.] If U is a solution of the martingale problem for A, then U r(t)= U (t) : { i Uei(t) < r defines a solution of the martingale problem for Ar. Uniqueness for } Ar follows by the argument outlined in Section A.6, and uniqueness for Ar implies uniqueness for A. Since uniqueness holds for the martingale problem for A, by Theorem A.15(c), uniqueness holds for C also. In general, if U is a solution of the martingale problem for A and i δUi(0) is a Poisson with mean measure yΛ, where Λ denotes , then (2.5) is a solution of the martingale problem forP C.

2.4. The genealogy and the number of ancestors. For each T > 0, there is a solution of (2.6) u˙ (t)= au (t)2 bu (t) T T − T satisfying uT (t) < for t < T and limt T uT (t)= . Every particle alive ∞ → − ∞ at time T is a descendent of some particle Ui(t) alive at time t < T satisfying Ui(t) < uT (t). Note that the converse is also true. If Ui(t) < uT (t), then Ui(t) has descendants alive at time T . In fact, a positive fraction of the particles alive at time T will be descendants of Ui(t). If Y (T ) > 0, then there are infinitely many particles alive at time T , but since ξ(t, [0, u (t))) < , T ∞ they are all descendants of finitely many ancestors alive at time t. Note that t ξ(t, [0, uT (t))) is nondecreasing and increases by jumps of +1. It is not→ possible to recover the full genealogy just from the levels since a new individual appearing at time t with level v could be the offspring of any existing individual with level Ui(t) < v. In Section 3, particles will be assigned a location (or type), and if these locations evolve in such a way that two particles have the same location only if one is the offspring of the other and then only at the instant of birth, it will be possible to reconstruct the full genealogy from the levels and locations.

2.5. Branching processes in random environments. Assume that a and b are functions of another ξ, say an irreducible, finite with generator Q. Then, for functions of the form f(l, u, n)= n f0(l)f1(u)= f0(l) i=1 g(ui), consider a scaled generator n Q r Arf(l, u, n)= rf1(u)Qf0(l)+ f(l, u, n) 2a(l) (g(v) 1) dv ui − Xi=1 Z n 2 g′(ui) + f(l, u, n) (a(l)ui √rb(l)ui) , − g(ui) Xi=1 10 T. G. KURTZ AND E. R. RODRIGUES which, as in (2.2), corresponds to a process with generator C f(l,n)= rQf(l,n)+ a(l)rn(f(l,n + 1) f(n)) r − + (ra(l) √rb(l))n(f(l,n 1) f(l,n)), b b − b − −b where f(l,n)= f (l)e λg n. The process corresponding to C is a branch- 0 − b b r ing process in a random environment determined by ξ. Writing the process correspondingb to Ar as

(ξ(rt),U1(t),...,UNr(t)) the process corresponding to Cr is (ξ(rt),Nr(t)). Note that in this example, the levels satisfy U˙ (t)= a(ξ(rt))U 2(t) √rb(ξ(rt))U (t). i i − i Let π be the stationary distribution for Q, and assume that l π(l)b(l)=0. Then, by Theorem 2.1 or [3], for example, P t Z(r)(t)= √r b(ξ(rs)) ds Z0 converges to a Z with variance parameter π(k)q (h (l) h (k))2 = 2 π(l)h (l)b(l) 2c, kl 0 − 0 − 0 ≡ Xk Xl Xl where h0(l) is a solution of Qh0(l)= b(l). In the limit, by Theorem 5.10 of [33] (applying a truncation argument to extend the boundedness assumption), the levels will satisfy 2 (2.7) dUi(t) = (aUi(t) + cUi(t)) dt + √2cUi(t) dW (t), where a = π(l)a(l). Applying ideas from [28], we can obtain convergence for the full system by consideringP the asymptotic behavior of the generator. Setting n g′(ui) h1(l, u, n)= h0(l)f1(u, n) ui , g(ui) Xi=1 we have 1 A f + h (l, u, n) r 1 √r 1   n r n 2 g′(ui) = f1(u, n) 2a(l) (g(v) 1) dv + f1(u, n) a(l)ui ui − g(ui) Xi=1 Z Xi=1 n n r 1 g′(ui) + h0(l)f1(u, n) ui 2a(l) (g(v) 1) dv √r g(ui) ! ui − Xi=1 Xi=1 Z POISSON REPRESENTATIONS 11

n 1 r + h0(l)f1(u, n) vg′(v) dv √r ui Xi=1 Z n 1 + h (l)f (u, n) (a(l)u2 √rb(l)u ) √r 0 1 j − j Xj=1 g′(ui)g′(uj) g′(uj)+ ujg′′(uj) ui + , × g(ui)g(uj ) g(uj )  i=j  X6 and passing to the limit as r , A (f + 1 h ) converges to →∞ r 1 √r 1

Af1(u, l)

∞ 2 g′(ui) e = f1(u) 2a(l) (g(v) 1) dv + f1(u) a(l)ui ui − g(ui) Xi Z Xi 2 g′(ui)g′(uj) ujg′(uj)+ uj g′′(uj) h0(l)b(l)f1(u) ujui + . − g(ui)g(uj ) g(uj ) j  i=j  X X6 1 Finally, we can find an additional perturbation h2 so that Ar(f1 + √r h1 + 1 r h2) converges to

∞ 2 g′(ui) Af1(u)= f1(u) 2a (g(v) 1) dv + f1(u) aui ui − g(ui) Xi Z Xi 2 g′(ui)g′(uj) ujg′(uj)+ uj g′′(uj) + cf1(u) ujui + . g(ui)g(uj ) g(uj ) j  i=j  X X6 This convergence assures convergence of the finite models to an infinite par- ticle model. The particle is the same as in Section 2.3, but the levels satisfy (2.7) where the Brownian motion W is the same for all levels. Let α and βg be as in Section 2.3, and note that

∞ ∞ β = (1 g(z)) dz = zg′(z) dz g − Z0 Z0 1 ∞ 2 = z g′′(z) dz. −2 Z0 We have from Lemma A.3

yβg ∞ ∞ αAf(y)= e− 2ay g(z) (g(v) 1) dv dz −  Z0 Zz ∞ 2 + y (az + cz)g′(z) dz Z0 12 T. G. KURTZ AND E. R. RODRIGUES

2 2 ∞ ∞ 2 + cy zg′(z) dz + cy z g′′(z) dz Z0  Z0  yβg 2 2 = e− ((ay + cy )β cyβ ) g − g = Cf(y), where b 2 Cf(y) = (ay + cy )f ′′(y)+ cyf ′(y), 1 which identifies the diffusion limit for r− Nr. Theorem A.15 can beb extended to coverb modelsb with non-Markovian en- vironments, that is, the process ξ is specified directly rather than through a generator. The diffusion limit is then obtained by verifying convergence of the level processes and applying Theorem A.12. For early work on diffusion approximations for branching processes in random environments, see [18, 26, 29] and also [9], Section 9.3.

3. Representations of measure-valued branching processes.

3.1. Branching Markov processes. We now consider particles with both a level ui and a location xi in a complete, separable metric E. Since the indexing of the particles is not important, we identify a state (x, u, n) of n our process with the counting measure µ(x,u) = i=1 δ(xi,ui). Let n P R log g dµ(x,u) f(x, u, n)= g(xi, ui)= e , Yi=1 where g : E [0, ) (0, 1]. We assume that as a function of x, g is in the domain (B×) of∞ the→ generator of a Markov process in E, g is continuously differentiableD in u and g(x, u)=1 for u r. We set ≥ n Bg(xi, ui) Arf(x, u, n)= f(x, u, n) g(xi, ui) Xi=1 n r (3.1) + f(x, u, n) 2a(xi) (g(xi, v) 1) dv ui − Xi=1 Z n 2 ∂ui g(xi, ui) + f(x, u, n) (a(xi)ui b(xi)ui) . − g(xi, ui) Xi=1 Each particle has a location Xi(t) in E and a level Ui(t) in [0, r]. The loca- tions evolve independently as Markov processes with generator B; the levels satisfy (3.2) U˙ (t)= a(X (t))U 2(t) b(X (t))U (t); i i i − i i POISSON REPRESENTATIONS 13 particles give birth at rates 2a(Xi(t))(r Ui(t)); the initial location of a new particle is the location of the parent at the− time of birth; and the initial level is uniformly distributed on [Ui(t), r]. Particles that reach level r die. Setting r n λg (xi) 1 n Pi=1 λg(xi) e− = g(xi)= r− 0 g(xi, z) dz and f(x,n)= i=1 g(xi)= e− , calculating as in Section 2.2, we have R Q b n n b b C f(x,n)= B f(x,n)+ ra(x )(f(b(x x ),n + 1) f(x,n)) r xi i | i − Xi=1 Xi=1 b n b b b + (ra(x ) b(x ))(f(d(x x ),n 1) f(x,n)), i − i | i − − Xi=1 b b where Bxi is the generator B applied to f(x,n) as a function of xi, b(x xi) is the collection of n + 1 particles in E obtained from x by adding a copy| of the ith particle xi, and d(x xi) is the collectionb of n 1 particles obtained | − n from x by deleting the ith particle, that is, if µx denotes i=1 δxi , then for z E, ∈ P n n

µb(x z) = δz + δxi , µd(x xj ) = δxi δxj . | | − Xi=1 Xi=1 If ra(z) b(z) 0 for all z E, then C is the generator of a branching Markov process− with≥ particle motion∈ determined by B, the birth rate for a particle at z E given by ra(z) and the death rate given by ra(z) b(z). With Theorem∈ A.15 in mind, we make the following assumptions− on B, a, b and r. C(E) is the space of continuous functions on E, C(E) the space of bounded continuous functions on E and M(E) the space of Borel measurable functions on E.

Condition 3.1. (i) B C(E) C(E), (B) is closed under multiplication and is separat- ing.⊂ × D n m (ii) f (A) satisfies f(x, u, n)= i=1 g(xi, ui), where g(z, v)= l=1(1 gl∈(z D)gl (v)) for gl (B), gl differentiable with support in [0, r] and− 1 2 1 ∈ D 2 Q Q 0 gl (z)gl (v) ρ < 1 for all l, z and v. ≤ 1 2 ≤ g (iii) There exists ψB C(E), ψB 0 and constants cg 0, for each g in (ii), such that ∈ ≥ ≥

sup Bg(x, u) cgψB(x), x E. u | |≤ ∈ 1 (iv) Defining B0 = (g, (ψB 1)− Bg) : g (B) , B0 is graph separable (see Section A.5{). ∨ ∈ D } (v) a, b M(E), a 0, r> 0 and ra b 0. ∈ ≥ − ≥ 14 T. G. KURTZ AND E. R. RODRIGUES

We have the following generalization of the results of Section 2.2.

Theorem 3.2. Assume Condition 3.1. For x En and u [0, )n, let ∈ ∈ ∞ n ui ψ(x, u)=1+ (ψ (x )+ a(x )+ b(x ) )e− B i i | i | Xi=1 and n r ψ(x)=1+ (ψ (x )+ a(x )+ b(x ) )(1 e− ). B i i | i | − Xi=1 If X is a solutione of the martingale problem for C satisfying t (3.3) E ψ(X(s)) ds < for all t 0, ∞ ≥ Z0  then there is a solution e(X, U) of the martingale problem for Ar such that X and X have the same distribution. e e Remarke 3.3. For many models, ψB , a and b will be uniformly bounded, and the moment conditions (3.3) will hold as long as E[X(0)] < . ∞ Proof of Theorem 3.2. Note that A f(x, u, n) (2r +(1+ r2 + r)d )erψ(x, u), | r |≤ g l l l lk where dg depends on the g1, g2, ∂vg2 and B g1 for all choices of l1,...,lj 1,...,m . The result then follows by application of Theorem A.15{ .  }⊂ { } Q Theorem 3.2 applies to finite branching Markov processes. Similar results also hold for locally finite processes.

Theorem 3.4. In addition to Condition 3.1, assume that

∞ g(x, u) 1 du + sup(u + u2) ∂ g(x, u) c ψ (x). | − | u ≤ g B Z0 u For x E and u [0, ) , let ∈ ∞ ∈ ∞ ∞ ∞ ui ψ(x, u)=1+ ψ (x )(1 + a(x )+ b(x ) )e− B i i | i | Xi=1 and

∞ r ψ(x)=1+ ψ (x )(1 + a(x )+ b(x ) )(1 e− ). B i i | i | − Xi=1 e POISSON REPRESENTATIONS 15

If X is a solution of the martingale problem for C satisfying t (3.4) E ψ(X(s)) ds < for all t 0, ∞ ≥ Z0  then there is a solution e(X, U) of the martingale problem for Ar such that X and X have the same distribution. e e Proof.e Note that A f(x, u, n) 2c erψ(x, u). r ≤ g The result then follows by application of Theorem A.15. 

Example 3.5. Suppose that a and b are bounded, and B is a diffusion operator with bounded drift and diffusion coefficients. Then we can take (B) to be the collection of nonnegative C2-functions with compact support D 1 and ψB(z)= (1+ z 2)β , for β> 0. If E[ i ψB(Xi(0))] < , then there exists a | | ∞ solution of the martingale problems for C satisfying sups t E[ i ψB(Xi(s))] < P ≤ and hence E[ t ψ(X(s)) ds] < . ∞ 0 ∞ P R 3.2. Basic limit theorem.e As in Section 2.3, if r , Af given by (3.1) becomes →∞ Bg(x , u ) Af(x, u)= f(x, u) i i g(xi, ui) Xi rg (3.5) + f(x, u) 2a(xi) (g(xi, v) 1) dv ui − Xi Z 2 ∂ui g(xi, ui) + f(x, u) (a(xi)ui b(xi)ui) , − g(xi, ui) Xi where g : E [0, ) (0, 1] has the property that there exists r such that × ∞ → g g(z, v)=1 for v > rg, and f(x, u)= i g(xi, ui). We can identify the state space of the corresponding process with a subset of (E [0, ))∞ or, since order is not important, with a subsetQ of (E [0, )),× the∞ counting mea- sures on E [0, ). Define N × ∞ × ∞ (E [0, )) = µ (E [0, )) : µ(E [0, u]) < 0 0 such that f(x, u)=0 for u uf . (See Section A.3 for a discussion of the appropriateR topologyR to use in≥ the infinite measure setting.) 16 T. G. KURTZ AND E. R. RODRIGUES

As r , the particle process converges to a process in which parti- →∞ cle locations evolve as independent Markov processes with generator B, levels satisfy (3.2), a particle with level Ui(t) gives birth to new particles at its location Xi(t) and level in the interval [Ui(t)+ c, Ui(t)+ d] at rate 2a(X (t))(d c). A particle dies when its level hits . The level of a parti- i − ∞ cle born at time t0 (or in the initial population, if t0 = 0) with initial level Ui(t0) satisfies

t Rt b(Xi(s)) ds Ui(t0)e− 0 Ui(t)= v , t Rt b(Xi(s)) ds 1 Ui(t0) e− 0 a(Xi(v)) dv − t0 until it hits infinity. If b 0 andRa is bounded away from zero, then U will ≤ i hit infinity in finite time. If we extend the path Xi back along its ancestral path to time zero, we would have

t R b(Xi(s)) ds u0e− 0 Ui(t) v , t R b(Xi(s)) ds ≥ 1 u e− 0 a(X (v)) dv − 0 0 i where u0 is the level of the particle’sR ancestor at time zero, and Xi(s) is the position of the ancestor at time s t. Since we are assuming that the ≤ initial position of an offspring is that of the parent, Xi is a solution of the martingale problem for B.

Proposition 3.6. Let (X,U) be a solution of the martingale problem for A given by (3.5). Let (Xr,U r) consist of the subset of particles for which Ui < r, that is,

δ r r = δ 1 (U (t)). (Xj (t),Uj (t)) (Xi(t),Ui(t)) [0,r) i X X If a(z)r b(z) 0 for all z E, then (Xr,U r) is a solution of the martingale − ≥ ∈ problem for Ar given by (3.1).

Remark 3.7. The condition a(z)r b(z) 0 for all z E ensures that − ≥ ∈ any particle that is above level r at time t will stay above level r at all future .

Proof of Proposition 3.6. The proposition follows by the observa- tion that Ar can be obtained from A by restricting the domain to f(x, u)= g(x , u ) for which g(z, v)=1 for v r.  i i i ≥ Q POISSON REPRESENTATIONS 17

3.3. The genealogy. Assume for the moment that a and b do not depend on x. If the location process has the property that at the time of a birth only the offspring has the same location as the parent (e.g., if the location process is Brownian motion), then the full genealogy can be recovered from knowledge of the levels and locations. The collection of ancestors at time t < T of the particles alive at time T is (X (t),U (t)) : U (t) < u (t) , where { i i i T } uT (t) is given by (2.6). The number of particles in this collection is non- decreasing in t and increases only by jumps of +1. The parent of the new particle is identifiable by the fact that only the parent and the offspring will be at the same location. If a and b depend on x, the full genealogy is still determined by the locations and levels of the particles, but recovering the genealogy is more complicated since it may not be possible to tell whether or not a particle (Xi(t),Ui(t)) has descendants alive at time T >t just from information avail- able at time t. However, some easy observations can be made. For example, if infx a(x) > 0 and supx b(x) < , then for t < T , all particles alive at time T are descendants of finitely many∞ particles alive at time t.

3.4. The measure-valued limit. The generator for a Dawson–Watanabe superprocess is typically of the form

Cf(µ)=exp h, µ ( Bh(y) F (h(y),y))µ(dy), {−h i} − − ZE b for f(µ)=exp h, µ , where h, µ = E h(y)µ(dy) and h is an appropriate function in ({−hB) (see,i} e.g., Theoremh i 9.4.3 of [9]). For arising D R fromb branching models with offspring distributions having finite variances, F should be of the form F (h(y),y)= a(y)h(y)2 + b(y)h(y). − For µ f (E), let α(µ,dx du) be the distribution of a Poisson ran- dom measure∈ M on E [0, ) with× mean measure µ Λ. Then setting h(y)= × ∞ × ∞(1 g(y, v)) dv, 0 − R f(µ)= αf(µ)= f(x, u)α(µ,dx du) × Z b ∞ = exp (g(y, v) 1) dv µ(dy) − ZE Z0  = exp h, µ . {−h i} Using Lemma A.3, we have

αAf(µ)= ∞ Bg(y, v) dv µ(dy)exp h, µ {−h i} ZE Z0 + ∞ 2a(y)g(y, z) ∞(g(y, v) 1) dv dz µ(dy)exp h, µ − {−h i} ZE Z0 Zz 18 T. G. KURTZ AND E. R. RODRIGUES

+ ∞(a(y)v2 b(y)v) ∂ g(y, v) dv µ(dy)exp h, µ − v {−h i} ZE Z0 = ∞ Bg(y, v) dv µ(dy)exp h, µ {−h i} ZE Z0 + ∞ 2a(y)g(y, z) ∞(g(y, v) 1) dv dz µ(dy)exp h, µ − {−h i} ZE Z0 Zz ∞(2a(y)v b(y))(g(y, v) 1) dv µ(dy)exp h, µ − − − {−h i} ZE Z0 = ∞ Bg(y, v) dv µ(dy)exp h, µ {−h i} ZE Z0 + ∞ 2a(y)g(y, z) ∞(g(y, v) 1) dv dz µ(dy)exp h, µ − {−h i} ZE Z0 Zz ∞ 2a(y) ∞(g(y, v) 1) dv µ(dy)exp h, µ − − {−h i} ZE Z0 Zz + ∞ b(y)(g(y, v) 1) dv µ(dy)exp h, µ − {−h i} ZE Z0 = ∞ Bg(y, v) dv µ(dy)exp h, µ {−h i} ZE Z0 2 + a(y) ∞(g(y, v) 1) dv µ(dy)exp h, µ − {−h i} ZE Z0  + ∞ b(y)(g(y, v) 1) dv µ(dy)exp h, µ − {−h i} ZE Z0 = ( Bh(y)+ a(y)h(y)2 b(y)h(y))µ(dy)exp h, µ = Cf(µ). − − {−h i} ZE But C is the generator for a superprocess, so for each µ f (E), thereb exists a solution Z of the martingale problem for C with Z(0)∈ M= µ and hence a solution (X,U) of the martingale problem for A with initial distribution α(µ, ). The· mapping γ : (E [0, )) (E) used in the application of The- Nf × ∞ → Mf orem A.15 is given by 1 lim δx , if the measures converge r r i γ δ(xi,ui) = →∞ ui r in the weak topology,  X≤  X   µ0, otherwise, where µ0 is a fixed element of f (E). The solution δ(X (t),U (t)) of the M i i martingale problem for A is a conditionally (see P POISSON REPRESENTATIONS 19

Section A.2) with Cox measure Z(t). Consequently, the particles determine Z by 1 (3.6) Z(t) = lim δX (t). r r i →∞ Ui(t) r X≤ r 1 Since by Proposition 3.6 and Theorem 3.2, Z (t)= r Ui(t) r δXi(t) is the normalized empirical measure for a branching Markov pro≤cess, (3.6) P gives the convergence of the normalized branching Markov process to the corresponding Dawson–Watanabe superprocess (cf. [4, 49]).

4. Examples and extensions.

4.1. A model with immigration. The simplest immigration process as- sumes that the space–time point process giving the arrival times and loca- tions of the immigrants is a Poisson process. Assuming temporal homogene- ity, immigration is introduced by adding the generator of a space–time-level Poisson random measure. Let ν be the intensity of immigration, that is, ν(A)∆t is approximately the probability that an individual immigrates into A E in a time interval of ∆t. The generator becomes ⊂ Bg(x , u ) Af(x, u)= f(x, u) i i g(xi, ui) Xi rg + f(x, u) 2a(x ) (g(x , v) 1) dv i i − i Zui (4.1) X rg + f(x, u) (g(z, v) 1)ν(dz) dv − Z0 ZE 2 ∂ui g(xi, ui) + f(x, u) (a(xi)ui b(xi)ui) . − g(xi, ui) Xi Noting that the generator for finite r is obtained from A by restricting the domain to the collection of g with rg r, if ra(x) b(x) 0, for all x, the generator of the corresponding branching≤ Markov process− ≥ with immigration is n n C f(x,n)= B f(x,n)+ ra(x )(f(b(x x ),n + 1) f(x,n)) r xi i | i − Xi=1 Xi=1 b n b b b + (ra(x ) b(x ))(f(d(x x ),n 1) f(x,n)) i − i | i − − Xi=1 b b + (f(b(x z),n + 1) f(x,n))ν(dz). | − ZE b b 20 T. G. KURTZ AND E. R. RODRIGUES

For r = , setting h(x)= 0∞(1 g(x, v)) dv as before, the generator for the measure-valued∞ process is − R Cf(µ)= αAf(µ) = ( Bh + ah2 bh, µ h, ν )exp h, µ , h− − i − h i {−h i} for f(µ)=exp h, µ . Earlyb results{−h on branchingi} Markov processes with immigration include [19, b27]. Work on limiting measure-valued processes with immigration in- cludes [8, 37–39, 48].

4.2. Conditioning on nonextinction. In the limiting model considered in Section 3.2, let a and b be constant and b< 0. Let τ be the time of extinction and let U (0) be the minimum of the initial levels. Then τ is the solution of ∗a bτ 1 U (0) (1 e− )=0, so − ∗ b − 1 U (0)a b τ = log ∗ − . − b U (0)a ∗ If Z(0) = µ0, then U (0) is exponentially distributed with parameter µ0(E) and ∗ b bµ (E) P τ > T = P U (0) < = 1 exp 0 . { } ∗ a(1 e bT ) − −a(1 e bT )  − −   − −  The case b = 0 is obtained by passing to the limit so that P τ > T = { } 1 exp µ0(E)/(aT ) . −As in{− Section 2.2.1,} conditioning on τ > T and letting T is equiv- alent to conditioning on the initial Poisson{ process} having a→∞ level at zero. The resulting generator becomes Bg(x , u ) Af(x, u)= f(x, u) i i g(xi, ui) Xi rg + f(x, u) 2a (g(xi, v) 1) dv ui − ui>0 Z (4.2) X rg + f(x, u)2a (g(x , v) 1) dv 0 − Z0 ∂ g(x , u ) + f(x, u) (au2 bu ) ui i i , i − i g(x , u ) u >0 i i Xi where the ui are the nonzero levels, and the generator for the conditioned measure-valued process is given by setting

α f(µ)= f(x, u)α (µ,dx du) 0 0 × Z 1 = g(z, 0)µ(dz)exp h, µ µ {−h i} | | ZE POISSON REPRESENTATIONS 21 and

α Af(µ)= Bh(y)+ ah(y)2 b(y)h(y),µ 0 h− − i 1 g(z, 0)µ(dz)exp h, µ × µ {−h i} | | ZE 1 + (Bg(z, 0) 2ag(z, 0)h(z))µ(dz)exp h, µ . µ − {−h i} | | ZE Note that α is the distribution of ξ = δ + ∞ δ , where U , i 0 0,Z0 i=1 (Ui,Zi) { i ≥ 1 is a Poisson process with intensity µ(E), and Z0,Z1,... are i.i.d. with distribution} µ( )/µ(E). P · For earlier work, see [10, 12, 20, 36, 40]. In particular, the particle at level zero in the construction above is the “immortal particle” of Evans [10].

4.3. Conditioning on extinction. As in Section 2.2.2, if a and b are con- stant and b> 0, then conditioning on extinction is equivalent to conditioning b b on U (0) > . Defining Vi(t)= Ui(t) , the generator for the conditioned ∗ a − a process is Bg(x , v ) Af(x, v)= f(x, v) i i g(xi, vi) Xi rg (4.3) + f(x, v) 2a (g(xi, v) 1) dv vi − Xi>0 Z 2 ∂vi g(xi, vi) + f(x, v) (avi bvi) , − g(xi, vi) Xi>0 and the generator of the measure-valued process is

Cf(µ)= ( Bh(y)+ ah(y)2 + bh(y))µ(dy)exp h, µ , − {−h i} ZE for f(µ)=expb h, µ . In other words, conditioning a supercritical process on extinction{−h replacesi} the supercritical process by a subcritical one. This resultb is originally due to Evans and O’Connell [11].

4.4. Models with multiple simultaneous births. We now consider continuous- time, branching Markov processes with general offspring distributions. The general theory of branching Markov processes was developed by Ikeda, Na- gasawa and Watanabe in a long series of papers [21–24] following earlier work by several authors. The particle representation is substantially more complicated and passage to the infinite population limit more delicate. 22 T. G. KURTZ AND E. R. RODRIGUES

As above, the particles move independently in E according to a generator B. A particle at position x E with level u gives birth to k offspring at rate (r) k (k 1)∈ (k + 1)ak (x)(r u) r− − . New particles have the location of the parent, but their levels− are uniformly distributed on [u, r). Then for f(x, u, n)= n i=1 g(xi, ui), Q Arf(x, u, n) n Bg(x , u ) = f(x, u, n) i i g(xi, ui) Xi=1 n (r) ∞ (k + 1)ak (xi) + f(x, u, n) k 1 r − Xi=1 Xk=1 k (4.4) g(xi, vl) 1 dv1 dvk k × [ui,r) " ! − # · · · Z Yl=1 n k+1 ∞ 2 (r) ui + f(x, u, n) r ak (xi) 1 − r Xi=1 Xk=1   u 1 + (k + 1) i − r 

∂ui g(xi, ui) b(xi)ui . − ! g(xi, ui) Now, the levels satisfy the equation k+1 ∞ U (t) U (t) U˙ (t)= r2a(r)(X (t)) 1 i 1 + (k + 1) i i k i − r − r k=1    (4.5) X b(X (t))U (t). − i i 1 r Defining g(x)= r 0 g(x, v) dv and integrating (4.4) with respect to α(n, du), the uniform measure on [0, r]n, we have that R

Arf(x, u, n)α(n, du) Z = Crf(x,n) n Bg(x ) (4.6) = f(x,n) i g(xi) Xi=1 n ∞ + f(x,n) ra(r)(x )[g(x )k 1] k i i − Xi=1 Xk=1 POISSON REPRESENTATIONS 23

n ∞ (r) 1 + f(x,n) r kak (xi) b(xi) 1 , − ! g(xi) − Xi=1 Xk=1   which is the generator of a branching process with multiple births with (r) (r) birth rates rak ( ), death rate r k∞=1 kak ( ) b( ) (provided this expression is nonnegative)· and particles moving according· − · to the generator B. The P (r) analog of Theorem 3.2 holds with a(xi) replaced by k(k + 1)ak (xi) in the definition of ψ and ψ. P We define e k ∞ u Λ(r)(x, u)= r(k + 1)a(r)(x) 1 1 k − − r Xk=1     and assume that lim Λ(r)(x, u) r →∞ k l ∞ (r) k u (4.7) = lim r(k + 1)a (x) ( 1)l+1 r k l →∞ " − r # Xk=1 Xl=1     Λ(x, u) ≡ exists uniformly for x E and u in bounded intervals. This condition is essentially (9.4.36) of [9∈]. Observe that u k+1 (r) ∞ 2 (r) u u Λ (x, v) dv = r ak (x) 1 1 + (k + 1) , 0 − r − r Z Xk=1    so that (4.5) becomes

Ui(t) (4.8) U˙ (t)= Λ(r)(X (s), v) dv b(X (t))U (t) i i − i i Z0 and r (r) ∞ 2 (r) Λ (x, v) dv = r kak (x), 0 Z Xk=1 1 r (r) so that the death rate for the branching process can be written as r− 0 Λ (x, v) dv b(x). R As− in [31], ∂m ∂mΛ(r)(x, u) Λ(r)(x, u) ≡ ∂um 24 T. G. KURTZ AND E. R. RODRIGUES

m+1 ∞ (r) (k + 1)k (k m + 1) (4.9) = ( 1) ak (x) · ·m · 1− − r − kX=m k m u − 1 . × − r   Each of the derivatives is a monotone function of u, and since m 1 (r) m 1 (r) ∂ − Λ (x, b) ∂ − Λ (x, a) − b = ∂mΛ(r)(x, u) du, Za it follows by the convergence of Λ(r) and induction on m that each ∂mΛ(r)(x, u) converges uniformly in u on bounded intervals that are bounded away from zero. Consequently, Λ is infinitely differentiable in u, and for 0 < u1 < u < , 2 ∞ lim sup ∂mΛ(r)(x, u) ∂mΛ(x, u) = 0. r u1 u u2| − | →∞ ≤ ≤ The fact that the derivatives alternate in sign implies that ∂1Λ(x, ) is com- pletely monotone and hence can be represented as ·

1 ∞ uz ∂ Λ(x, u)= e− ν(x, dz) Z0 for some σ-finite measure ν(x, ). Writing νb(x, ) = 2a0(x)δ0 + ν(x, ) with ν(x, 0 )=0, · · · { } b ∞ 1 b uz Λ(x, u) = 2a (x)u + z− (1 e− )ν(x, dz). 0 − Z0 Let g satisfy g(x, v)=1, for v u , and define ≥ g ug h(x, u)= (1 g(x, v)) dv. − Zu If r > ug and there are n particles below level r, then (4.4) may be written as

Arf(x, u, n) n Bg(x , u ) = f(x, u, n) i i g(xi, ui) Xi=1 + f(x, u, n) n ∞ r(k + 1) × Xi=1 Xk=1 POISSON REPRESENTATIONS 25

u + h(x , u ) k u k a(r)(x ) 1 i i i 1 i × k i − r − − r      n ui (r) ∂ui g(xi, ui) + f(x, u, n) Λ (xi, v) dv b(xi)ui . 0 − g(xi, ui) Xi=1 Z  Then, by (4.7) and the definition of h, we have Af(x, u)

lim Arf(x, u) r ≡ →∞ Bg(x , u ) = f(x, u) i i g(xi, ui) Xi + f(x, u) (Λ(x , u ) Λ(x , u + h(x , u ))) i i − i i i i Xi ui ∂ui g(xi, ui) + f(x, u) Λ(xi, v) dv b(xi)ui 0 − g(xi, ui) Xi Z  Bg(x , u ) = f(x, u) i i g(xi, ui) Xi ∞ + f(x, u) 2a0(xi) (g(xi, v) 1) dv ui − Xi Z ∞ ∞ z Ru (g(xi,v) 1) dv 1 zui + f(x, u) (e i − 1)z− e− ν(xi, dz) 0 − Xi Z + f(x, u) a (x )u2 b(x )u 0 i i − i i Xi 

∞ 1 1 uiz ∂ui g(xi, ui) + z− (u z− (1 e− ))ν(x , dz) . i − − i g(x , u ) Z0  i i Note that the second term on the right-hand side has the same interpretation as the second term on the right-hand side of (3.5). To understand the third term, recall that if ξ = i δτi is a Poisson process on [0, ) with parameter λ, then ∞ P ∞ λ R (g(v) 1) dv E g(τi) = e 0 − . h i Consequently, the third termY determines bursts of simultaneous offspring at the location xi of the parent and with levels forming a Poisson process with intensity z on [u , ). i ∞ 26 T. G. KURTZ AND E. R. RODRIGUES

Setting h(x)= h(x, 0) = ∞(1 g(x, v)) dv and f(µ)=exp h, µ , 0 − {−h i} R Cf(µ)= αAf(µ) b h(y) b = Bh(y)+ Λ(y, z) dz b(y)h(y) µ(dy)exp h, µ . − − {−h i} ZE Z0  Based on the above calculations, we have the following theorem.

Theorem 4.1. Let B C(E) C(E) satisfy Condition 3.1, and let the ⊂ martingale problem for B be well posed.× Assume that for r r , ≥ 0 ∞ inf r ka(r)(x) b(x) 0 x k − ≥  Xk=1  and that the convergence in (4.7) is uniform in x. Let K(0) be a finite random measure on E, and let ξr be a solution of the martingale problem for r Ar such that ξ (0) is conditionally Poisson on E [0, r] with mean measure K(0) Λ. Then ξr ξ where ξ is a solution of the× martingale problem for × ⇒ A.

Proof. For r>q, let ξ(q),r denote the restriction of ξr to E [0,q] and (q) (q),r (q) × similarly for ξ . It is enough to prove that ξ ξ for each q > r0. The (q),r ⇒ generator for ξ is the restriction of Ar to functions f (Ar) such that the corresponding g satisfies g(x, u)=1 for u q. For f∈ Dof this form, by ≥ (4.9), Ar,qf = Arf satisfies

Ar,qf(x, u, n) n Bg(x , u ) = f(x, u, n) i i g(xi, ui) Xi=1 n ∞ 1 + f(x, u, n) ( 1)m+1∂mΛ(r)(x ,q) m! − i m=1 Xi=1 X m g(xi, vl) 1 dv1 dvm m × [ui,q) " ! − # · · · Z Yl=1 n ui (r) ∂ui g(xi, ui) + f(x, u, n) Λ (x, v) dv b(xi)ui , 0 − g(xi, ui) Xi=1 Z  and the corresponding branching Markov process has generator

Cr,qf(x,n) POISSON REPRESENTATIONS 27

n Bg(x ) = f(x,n) i g(xi) Xi=1 n ∞ 1 + f(x,n) ( 1)m+1∂mΛ(r)(x ,q)qm[g(x )m 1] (m + 1)! − i i − Xi=1 mX=1 n q 1 (r) 1 + f(x,n) Λ (xi, v) dv b(xi) 1 . q 0 − g(xi) − Xi=1  Z   The convergence of ξ(q),r follows by the convergence assumptions on Λ(r). 

The measure ν(x, ) is nonzero only if the offspring distribution has a (r) · “heavy tail.” If ak (x)= ak(x) and

∞ (k + 1)ka (x) < , k ∞ Xk=1 then k ∞ u ∞ Λ(x, u) = lim r(k + 1)ak(x) 1 1 = (k + 1)kak(x)u r →∞ − − r Xk=1     Xk=1 and Bg(x , u ) Af(x, u)= f(x, u) i i g(xi, ui) Xi ∞ ug + f(x, u) (k + 1)kak(xi) [g(xi, v) 1] dv ui − Xi Xk=1 Z ∞ (k + 1)kak(xi) 2 ∂ui g(xi, ui) + f(x, u) ui b(xi)ui , 2 − ! g(xi, ui) Xi Xk=1 which is essentially (3.5). For scalar branching processes with general offspring distributions, con- vergence to possibly discontinuous continuous state branching processes was proved by Grimvall [16] (see [9], Section 9.1). Convergence in the measure- valued setting is given in [49] and [4] for offspring distributions with finite second moment and more generally in [9], Theorem 9.4.3. Fitzsimmons [13] gives a very general construction of these processes. If ν is not zero, then the genealogy of the process is much more compli- cated than that described in Sections 2.4 and 3.3. Assume that Λ and b do 28 T. G. KURTZ AND E. R. RODRIGUES not depend on x, and define u 2 ∞ 2 uz Λ(u) Λ(v) dv bu = 2a u bu + z− (zu 1+ e− )ν(dz). ≡ − 0 − − Z0 Z0 If ubT satisfies u˙ T (t)= Λ(uT (t)), uT (t) < for t < T and limt T uT (t)= , then it is still the case that the collection∞ of ancestors at→ time− tb < T ∞of the population alive at time T is (X (t),U (t)) : U (t) < u (t) , but u may not exist. In fact, since if { i i i T } T u˙ = Λ(u) u(t2) 1 b dv = t t , 2 − 1 Zu(t1) Λ(v) uT exists if and only if b 1 ∞ dv < ∞ Zu Λ(v) for u sufficiently large, which always holds if a0 > 0. In the critical and subcritical cases, this condition isb equivalent to extinction with probability one as was noted by Bertoin and Le Gall ([2], page 167). This finite ancestry property or coming down from infinity of the geneal- ogy has been studied for a variety of population models. See [47] and [1] for results in the Fleming–Viot setting. The equivalence of the conditions for Fleming–Viot and Dawson–Watanabe processes is given in ([2], page 171). The argument in Section 2.2.3 can undoubtedly be extended to the present setting. This development will be carried out elsewhere.

4.5. Model with exponentially distributed levels. The discrete models that we have considered have been formulated with levels that are uniformly dis- tributed on an interval. That is not necessary, and other distributions may be convenient in other contexts. We illustrate this flexibility by formulating a model for a simple branching process with levels that are exponentially distributed. The dynamics of the levels change, and the correct dynamics are determined by essentially working backwards from the answer. As before, let f(u, n)= n g(u ) where 0 g 1 and g(u )=1 for i=1 i ≤ ≤ i ui ug. Let ≥ Q n n ∞ v/r g′(ui) Arf(u, n)= f(u, n) 2a e− (g(v) 1) dv + f(u, n) Gr(ui) , ui − g(ui) Xi=1 Z Xi=1 where Gr will be determined below. Note that a particle at level ui is giving ui/r birth at rate 2are− , and the levels satisfy

U˙ i(t)= Gr(Ui(t)). POISSON REPRESENTATIONS 29

Let αr(n, du) be the distribution of n independent exponential random vari- λg 1 v/r ables with mean r, and define e− = r− 0∞ g(v)e− dv so

R λg n f(n)= f(u, n)αr(n, du)= e− . Z To calculate Arf(u,b n)αr(n, du), observe that

R 1 ∞ z/r ∞ v/r r− 2a e− g(z) e− (g(v) 1) dv dz − Z0 Zz

2λg ∞ 2z/r = are− 2a e− g(z) dz, − Z0 and assuming Gr(0) = 0,

1 ∞ z/r r− e− Gr(z)g′(z) dz Z0 1 ∞ z/r 1 = r− e− (G′ (z) r− G (z))(g(z) 1) dz − r − r − Z0 1 ∞ z/r 1 = r− e− (G′ (z) r− G (z))g(z) dz − r − r Z0 1 ∞ z/r 1 + r− e− (G′ (z) r− G (z)) dz. r − r Z0 Then for

1 z/r d z/r z/r G′ (z) r− G (z)= e (e− G (z)) = 2ar(1 e− ) b, r − r dz r − − we have

z/r z/r r 2z/r z/r e− G (z) = 2ar r(1 e− ) (1 e− ) br(1 e− ) r − − 2 − − −   and

Arf(u, n)αr(n, du) Z

λg(n 1) 2λg 1 ∞ z/r 1 = ne− − are− r− e− (G′ (z) r− G (z) − r − r  Z0 z/r + 2are− )g(z) dz

1 ∞ z/r 1 + r− e− (G′ (z) r− G (z)) dz r − r Z0  = Crf(n),

b 30 T. G. KURTZ AND E. R. RODRIGUES where

(4.10) C f(n)= ran(f(n + 1) f(n)) + (ra b)n(f(n 1) f(n)) r − − − − is the generatorb of a branchingb process.b b b Note that as r , G (z) converges to az2 bz, and hence, A converges →∞ r − r to A given by (2.4).

4.6. Multitype branching processes. We now consider a branching parti- cle system with m possible types, S = 1, 2,...,m . We assume that a parti- { } cle of type ζ S gives birth to a particle of type ζ S at rate ra(r)(ζ ,ζ ) 1 ∈ 2 ∈ 1 2 and dies at rate ra(r)(ζ ) b(r)(ζ ), where a(r)(ζ )= a(r)(ζ , j). 1 − 1 1 j S 1 The fact that the ordered representations constructed∈ for the previous examples give the correct measure-valued processes dependP s on the fact that observing a birth event in the measure-valued process gives no information about the levels of the particles after the birth event. That, in turn, depends on the offspring being indistinguishable from the parent. Since in the current model, the type of an offspring may differ from the type of the parent, we need to find a way to “preserve ignorance” about the levels when the type of the offspring is different. We accomplish this goal by randomizing the assignment of the parent and offspring to the original level of the parent and a new level. Let f(ζ, u, n) be of the form

n f(ζ, u, n)= g(ζi, ui). Yi=1 Then the generator of the ordered representation of the branching process described above is given by

Arf(ζ, u, n) n (r) = f(ζ, u, n) 2a (ζi, j) i=1 j S X X∈ r 1 g(ζ , u )g(j, v)+ g(ζ , v)g(j, u ) i i i i 1 dv × 2 g(ζ , u ) − Zui   i i   n (r) 2 (r) ∂ui g(ζi, ui) + f(ζ, u, n) [a (ζi)ui b (ζi)ui] , − g(ζi, ui) Xi=1 where as before, each level satisfies d U (r)(t)= a(r)(X (t))U 2(t) b(r)(X (t))U (t). dt i i i − i i POISSON REPRESENTATIONS 31

(r) (r) Let Q h(ζ)= j S a (ζ, j)[h(j) h(ζ)]. Because of the randomization of the level assignments∈ at each birth− event, it follows that P t h(X (t)) (r U (s))Q(r)h(X (s)) ds i − − i i Zτi is a martingale. Taking αr(n, du) as before, we have that

Arf(ζ, u, n)α(n, du)= Crf(ζ,n) Z n = f(ζ,n) ra(r)(ζ , j)[g(j) 1] i − i=1 j S X X∈ n (r) (r) 1 + f(ζ,n) [ra (ζi) b (ζi)] 1 , − g(ζi) − Xi=1   n 1 r where f(ζ,n)= i=1 g(ζi) and g(ζi)= r 0 g(ζi, v) dv. Hence, Crf(ζ,n) is the generator of a multitype branching process. Assume that Q R a(ζ, j) = lim a(r)(ζ, j), r →∞ a(ζ) = lim a(r)(ζ)= a(ζ, j), r →∞ j S X∈ b(ζ) = lim b(r)(ζ) r →∞ and that Qh(ζ)= a(ζ, j)[h(j) h(ζ)] − j S X∈ is the generator of an irreducible, finite state Markov chain. Let π denote the unique stationary distribution for Q. It is clear from the of the Markov chain that in the limit, the levels must satisfy d U (t)= aU 2(t) bU (t), dt i i − i where a = j π(j)a(j) and b = j π(j)b(j). We can make this observation precise by analyzing the asymptotic behav- P P 1 ior of the generator. Taking g(ζ, u) = exp( h0(u)+ r h(ζ, u)), where h(ζ, u) and h (u) are equal to zero if u u , and− letting r , we have that 0 ≥ g →∞ 1 lim f(ζ, u) = lim exp h0(ui)+ h(ζi, ui) r r →∞ →∞ − r  Xi Xi  32 T. G. KURTZ AND E. R. RODRIGUES

= exp h (u ) f(u) − 0 i ≡  Xi  and since

g(ζi, ui)g(j, v)+ g(ζi, v)g(j, ui) g(ζi, ui) −1 −1 h0(v) r h(j,v) r (h(ζi,v)+h(j,ui) h(ζi,ui)) = e− (e + e − ),

lim Arf(ζ, u) r →∞ ug h0(v) = f(u) 2a(ζ ) [e− 1] dv i − i  Zui (4.11) X + a(ζ , j)[h(j, u ) h(ζ , u )] i i − i i j S X∈ [a(ζ )u2 b(ζ )u ] ∂ h (u ) . − i i − i i ui 0 i  If π(j)G(j, u) 0 for all u, then there exists h such that ≡ P a(ζ, j)[h(j, u) h(ζ, u)] = G(ζ, u). − j S X∈ Consequently, there exists h such that the right-hand side of (4.11) becomes

ug h0(v) 2 Af(u)= f(u) 2a [e− 1] dv [aui bui] ∂ui h0(ui) , ui − − − Xi  Z  which is just a rewriting of (2.4), and hence we have convergence of the normalized total population to the Feller diffusion. For earlier work, see [14, 25, 29] and Section 9.2 of [9].

4.7. Models with catastrophic death. Now consider n Bg(xi, ui) Arf(x, u, n)= f(x, u, n) g(xi, ui) Xi=1 n r + f(x, u, n) 2a(x ) (g(x , v) 1) dv i i − i=1 Zui (4.12) X n 2 ∂ui g(xi, ui) + f(x, u, n) (a(xi)ui b(xi)ui) − g(xi, ui) Xi=1 + (f(x, c(u, x, v),n) f(x, u, n))γ(dv), − ZV POISSON REPRESENTATIONS 33 where γ is a σ-finite measure on a measurable space (V, ), V c(u, x, v) = (u1ρ(x1, v), u2ρ(x2, v),...) and ρ(x , v) 1. Then as in Section 3.1 i ≥ n n C f(x,n)= B f(x,n)+ ra(x )(f(b(x x ),n + 1) f(x,n)) r xi i | i − Xi=1 Xi=1 b n b b b + (ra(x ) b(x ))(f(d(x x ),n 1) f(x,n)) i − i | i − − Xi=1 n b b 1 1 + (ρ− (xi, v)g(xi) + (1 ρ− (xi, v))) f(x,n) γ(dv). V − − ! Z Yi=1 For simplicity, assume that γ(V )b< . Then at rate γ(V ) anb event occurs in which an element v is selected from∞ V , and given v, particles are inde- pendently killed, with the probability that a particle at xi survives being 1 ρ− (xi, v). Letting r to obtain A and integrating, →∞ αAf = ( Bh(y)+ a(y)h(y)2 b(y)h(y))µ(dy)exp h, µ − − {−h i} ZE 1 + (exp ρ− ( , v)h, µ exp h, µ )γ(dv) {−h · i} − {−h i} ZV = Cf(µ). Branching processes with catastrophes have been considered in a series of b papers by Pakes [41–46] and by Grey [15].

APPENDIX A.1. Poisson random measures. Let (S, ) be a measurable space, and let ν be a σ-finite measure on . ξ is a PoissonS random measure with mean measure ν if: S (a) ξ is a random counting measure on S; (b) for each A with ν(A) < , ξ(A) is Poisson distributed with param- eter ν(A); ∈ S ∞ (c) for A , A ,... disjoint, ξ(A ),ξ(A ),... are independent. 1 2 ∈ S 1 2 Lemma A.1. If H : S S is Borel measurable and ξ(A)= ξ(H 1(A)), → 0 − then ξ is a Poisson random measure on S0 with mean measure ν given by 1 ν(A)= ν(H− (A)). b b b b 34 T. G. KURTZ AND E. R. RODRIGUES

Remark A.2. ν need not be σ-finite even if ν is, but the meaning of the lemma should still be clear. σ-finite or not, ν(A)= if and only if ∞ ξ(A)= a.s. ∞ b b b Proof of Lemma A.1. The lemma follows from the fact that A1, A2,... 1 1 disjoint implies H− (A1),H− (A2),... are disjoint. 

Lemma A.3. If ξ is a Poisson random measure with mean measure ν and f L1(ν), then ∈ R f(z)ξ(dz) R (ef 1) dν (A.1) E[e ]= e − ,

(A.2) E f(z)ξ(dz) = f dν, Var f(z)ξ(dz) = f 2 dν, Z  Z Z  Z allowing = . Letting∞ξ = ∞ δ , for g 0 with log g L1(ν), i Zi ≥ ∈ P R (g 1) dν E g(Zi) = e − . Yi  Similarly, if hg, g 1 L1(ν), then − ∈ R (g 1) dν E h(Zj ) g(Zi) = hg dν e − Xj Yi  Z and 2 R (g 1) dν E h(Zi)h(Zj) g(Zk) = hg dν e − .  i=j k  Z  X6 Y Proof. The independence properties of ξ imply (A.1) for simple func- tions. The general case follows by approximation. The other identities follow in a similar manner. Note that the integrability of the random variables in the expectations above can be verified by replacing g by ( g a)1 + 1 c | |∧ A A and h by ( h a)1A for 0

ξ = δ(Xi,Ui) Xi is a Poisson random measure on [0, )2 with mean measure λµ Λ, were ∞ X × µX is the law of X1. POISSON REPRESENTATIONS 35

If κ = E[ 1 ] < , then Xi ∞

ξ = δXiUi X is a Poisson random measure onb [0, ) with mean measure λκΛ. ∞ Proof. By Lemma A.1, ξ is a Poisson random measure with mean measure given by b ∞ 1 1 ν[0, c]= λµ Λ (x, u) : xu c = λ P X− uc− du X × { ≤ } { ≥ } Z0 1  b = λcE[X− ]. A.2. Conditionally Poisson systems. We begin by considering general conditionally Poisson systems or Cox processes. Consider (S, d) a metric space, and let ξ be a random counting measure on S and Ξ be a locally finite random measure on S. [A measure ν on S is locally finite if for each x S, there exists an ǫ> 0 such that ν(Bε(x)) < .] We say that ξ is conditionally∈ Poisson with Cox measure Ξ if, conditioned∞ on Ξ, ξ is a Poisson random measure with mean measure Ξ. This requirement is equivalent to R f dξ R (1 e−f ) dΞ E[e− S ]= E[e− S − ], for all nonnegative f M(S), where M(S) is the set of all Borel measurable ∈ R f dµ functions on S. Since the collection of functions Ff (µ)= e− S is closed under multiplication and separates points in the space of locally finite mea- sures, the distribution of Ξ determines the distribution of ξ. We are actually interested in the conditionally Poisson system on S [0, ) with Cox measure Ξ Λ, where Λ is Lebesgue measure. Then for× nonnegative∞ f M(S), we have× ∈ R f dξ K R (1 e−f ) dΞ E[e− S×[0,K] ]= E[e− S − ], and the distribution of ξ determines the distribution of Ξ, where we con- sider f M(S) to be a function on S [0, K] satisfying f(x, u)= f(x). In particular,∈ × 1 Ξ(f) = lim f dξ a.s. K K S [0,K] →∞ Z ×

Lemma A.5. Suppose ξ is a conditionally Poisson random measure on S [0, ) with Cox measure Ξ Λ, and let f M(S), 0 f 1. Then for C,D>× ∞0, × ∈ ≤ ≤ KD (A.3) P f dξ C + P f dΞ D S [0,K] ≥ ≤ C S ≥ Z ×  Z  36 T. G. KURTZ AND E. R. RODRIGUES and C−1 R f dξ E[1 e− S×[0,K] ] (A.4) P f dΞ C − −C−1 . S ≥ ≤ 1 e Ke Z  − − Let ξ , α be a collection of conditionally Poisson random measures { α ∈ A} on S [0, ) with Cox measures Ξα Λ, and let f M(S), 0 f 1. Then × f∞ dξ , α is stochastically× bounded if and∈ only if ≤ f≤ dΞ , α { S [0,K] α ∈ A} { S α ∈ ×is stochastically bounded. A}R R

Proof. Since E[ S [0,K] f dξ Ξ]= K S f dΞ, × | R K f dRΞ KD P f dξ C Ξ S 1 + 1 , RS f dΞ D S [0,K] ≥ ≤ C ∧ ≤ C { ≥ } Z ×  R and taking expectations gives (A.3). By (A.1)

R εf dξ K R (1 e−εf ) dΞ E[1 e− S×[0,K] ]= E[1 e− S − ] − − εKe−ε R f dΞ E[1 e− S ] ≥ − εKe−εC (1 e− )P f dΞ C , ≥ − ≥ ZS  1 and taking ε = C− gives (A.4). The final statement follows from the two inequalities. 

Let ξ = δ be a point process on S, and let U be independent ran- Xi { i} dom variables, uniformly distributed on [0, r] and independent of ξ. Define b P (A.5) ξ = δ , Ξ = r 1ξ. (Xi,Ui) r − b Then for f 0 on S [0, rX], ≥ × b r r RS× ,r f dξ 1 f(Xi,u) R F (x)Ξr(dx) (A.6) E[e− [0 ] Ξr]= r− e− du = e− S f , | 0 Yi  Z  where r r r 1 f(x,u) 1 f(x,u) F (x)= r log e− du = r log 1 (1 e− ) du . f − r − − r − Z0  Z0  We have the following analog of Lemma A.5.

Lemma A.6. Suppose ξ and Ξr are given by (A.5), and let f M(S), 0 f 1. Then for C,D> 0 and K r, ∈ ≤ ≤ ≤ KD (A.7) P f dξ C + P f dΞr D S [0,K] ≥ ≤ C S ≥ Z ×  Z  POISSON REPRESENTATIONS 37 and C−1 R f dξ E[1 e− S×[0,K] ] (A.8) P f dΞ C . r − −C−1 S ≥ ≤ 1 e Ke Z  − −

Proof. Since E[ S [0,K] f dξ Ξr]= K S f dΞr, × | R K f dΞR S r KD 1 P f dξ C Ξr 1 + R f dΞr D , S [0,K] ≥ ≤ C ∧ ≤ C { S ≥ } Z ×  R and taking expectations gives (A.7). Defining

r K εf(x) ε G (x)= r log 1 (1 e− ) εKe− f(x), K,ε,f − − r − ≥   by (A.6)

r R εf dξ R G dΞr E[1 e− S×[0,K] ]= E[1 e− S K,ε,f ] − − −ε εKe R f dΞr E[1 e− S ] ≥ − εKe−εC (1 e− )P f dΞ C , ≥ − r ≥ ZS  1 and taking ε = C− gives (A.8). 

Lemma A.7. Suppose ξ is a conditionally Poisson random measure on S [0, ) with Cox measure Ξ Λ. If Ξ(S) < a.s., then we can write × ∞ × ∞ ξ = ∞ δ with U < U < a.s. and X exchangeable. i=1 (Xi,Ui) 1 2 · · · { i}

P Ξ Proof. Let Xi be exchangeable with de Finetti measure Ξ , and let { } | | Y be a unit Poisson process with jump times Si independent of of Xi e { } { } and Ξ. Define ξ = ∞ δ −1 , and note that i=1 (Xi, Ξ Si) e | | e P −1 R fdξ f(Xi, Ξ Si) E[e− e]=e E e− e | | Yi  −1 f(z, Ξ Si) 1 = E e− | | Ξ − Ξ(dz) | | Yi Z  ∞ f(z, Ξ −1s) 1 = E exp 1 e− | | Ξ − Ξ(dz) ds − − | |   Z0  Z   ∞ f(z,u) = E exp (1 e− )Ξ(dz) du . − −   Z0 Z  38 T. G. KURTZ AND E. R. RODRIGUES

Consequently, ξ is conditionally Poisson with Cox measure Ξ Λ, and ξ and ξ have the same distribution.  × e e As in Lemma A.4, we have the following.

Lemma A.8. Suppose ξ is a conditionally Poisson random measure on S [0, )2 with Cox measure Ξ Λ, where Ξ is a random measure on S ×[0, ∞). Suppose × × ∞ 1 Ξ(A)= 1A(x)Ξ(dx dy) S [0, ) y × Z × ∞ defines a locally finiteb random measure on S. Then writing ξ = i δ(Xi,Yi,Ui), ξ = δ is a conditionally Poisson random measure on S [0, ) i (Xi,YiUi) P × ∞ with Cox measure Ξ Λ, and hence b P × 1 1 y− f(bx)Ξ(dx dy)= Ξ(f) = lim f dξ a.s. S [0, ) × K K S [0,K] Z × ∞ →∞ Z × b b A.3. Convergence results. Let hk, k = 1, 2,... C(S) satisfy 0 hk 1 and x : h (x) > 0 = S, where{ C(S) denotes}⊂ the space of bounded≤ ≤ k{ k } continuous functions on S, and let h (S) be the collection of Borel mea- M{ k} sures onSS satisfying h dν < , for all k, topologized by the requirement S k ∞ that ν ν if and only if fh dν fh dν for all f C(S) and k; n → R S k n → S k ∈ that is, the measures dνk = h dν converge weakly for each k. Similarly, n R k n R let h (S [0, )) be the space of Borel measures on S [0, ) satis- M{ k} × ∞ × ∞ fying h dµ < for all k = 1, 2,... and K> 0, topologized by the S [0,K] k ∞ requirement× that µ µ if and only if R n →

fhk dµn fhk dµ, S [0, ) → S [0, ) Z × ∞ Z × ∞ for all k and f C(S [0, )) such that the support of f is contained in ∈ × ∞ S [0, K] for some K> 0. Note that in both cases, hk is metrizable. To simplify× notation, let M{ }

h (S)= f C(S) : f chk for some c> 0 and hk . C{ k} { ∈ | |≤ } Then convergence in h (S) is equivalent to convergence of f dνn for M{ k} S all f h (S). ∈C{ k} R Theorem A.9. Let ξn be a sequence of conditionally Poisson ran- dom measures on S [0,{ )}with Cox measures Ξn Λ . Then ξn ξ in × ∞ n { × } ⇒ hk (S [0, )) if and only if Ξ Ξ in hk (S). If the limit holds, thenM{ ξ} is conditionally× ∞ Poisson with Cox⇒ measureM{ Ξ} Λ. × POISSON REPRESENTATIONS 39

n Proof. Suppose ξ ξ in hk (S [0, )). Then for each f C(S), f 0, each k, and all but⇒ countablyM{ many} × K ∞ ∈ ≥ R fh dξ R fh dξn E[e− S×[0,K] k ] = lim E[e− S×[0,K] k ] n →∞ K R h−1(1 e−fhk )h dΞn = lim E[e− S k − k ]. n →∞ For g 0 and K satisfying sup K 1g(x)h (x) < 1, let ≥ x − k 1 1 h− (x) log(1 K g(x)h (x)), h (x) > 0, f(x)= − k − − k k K 1g(x), h (x) = 0,  − k and we see that

n R gh dΞ R fhk dξ lim E[e− S k ]= E[e− S×[0,K] ] n →∞ n n exists. Since ξ ξ in hk (S [0, )), S [0,K] hk dξ is stochastically ⇒ M{ } × ∞n { × } bounded and by Lemma A.5, hk dΞ mustR be stochastically bounded. Tightness follows similarly. Consequently,{ } Ξn is relatively compact in R { } hk (S) in distribution, and the unique limit Ξ is determined by the fact Mthat{ }

R gh dΞ R fhk dξ E[e− S k ]= E[e− S×[0,K] ], for g and f related as above. The proof of the converse is similar. 

n Theorem A.10. For each n = 1, 2,..., let rn > 0 and ξ be a point process on S [0, r ], and define × n n 1 n (A.9) Ξ (dx)= ξ (dx [0, rn]). rn ×

n n R f(x,u)ξn(dx du) R F (x)Ξ (dx) Suppose for f 0, E[e ]= E[e− f ], where ≥ − × rn rn n 1 f(x,u) 1 f(x,u) F (x)= r log e− du = r log 1 (1 e− ) du , f − n r − n − r − n Z0  n Z0  that is, the [0, rn] components are independent, uniformly distributed, and n n independent of Ξ . Then assuming rn , ξ ξ in hk (S [0, )) if n →∞ ⇒ M{ } × ∞ and only if Ξ Ξ in hk (S). If the limit holds, then ξ is conditionally Poisson with Cox⇒ measureM{ Ξ} Λ. × Proof. For g ,f 0, g C ([0, )), f C(S) and f(x, u)= h (x) 0 0 ≥ 0 ∈ c ∞ 0 ∈ k × f (x)g (u), F n(x) ∞(1 e f(x,u)) du, and the remainder of the proof is 0 0 f → 0 − − similar to that of Theorem A.9.  R 40 T. G. KURTZ AND E. R. RODRIGUES

These convergence theorems apply only to the one-dimensional distribu- tions of the models considered in this paper. To address convergence as pro- 1 cesses, note that for finite r and Ξr(t,dx)= r− ξ(t,dx [0, r]), the models satisfy ×

r R f(x,u)ξ(t,dx du) Ξr R F (x)Ξr (t,dx) (A.10) E[e− S×[0,r] × ]= e− f , |Ft where r r r 1 f(x,u) 1 f(x,u) F (x)= r log e− du = r log 1 (1 e− ) du , f − r − − r − Z0  Z0  and the r = models satisfy ∞ R f dξ(t) Ξ K R (1 e−f ) dΞ(t) (A.11) E[e− S×[0,K] ]= e− S − , |Ft for f hk (S). The following estimates imply that convergence of the finite-dimensional∈C{ } distributions for ξn imply convergence of the finite-dimen- n n sional distributions for Ξ (or Ξrn , assuming rn ); however, convergence of the finite-dimensional distributions of Ξn may→∞ not imply convergence of the finite-dimensional distributions of of ξn.

Lemma A.11. Suppose ξ is a conditionally Poisson random measure on S [0, ) with Cox measure Ξ Λ, Ξ with values in h (S). Then for × ∞ × M{ k} each f h (S) and δ, K, K′ > 0, ∈C{ k}

1 P K− f dξ f dΞ δ − ≥  ZS [0,K] ZS  × C (A.12) + P f 2 dΞ >C ≤ Kδ2 Z  −1 2 C R ′ f dξ C E[1 e− S×[0,K ] ] + − −1 . ≤ Kδ2 1 e K′e−C − − Suppose ξ satisfies (A.10). Then for each f h (S), δ> 0 and 0 < K, ∈C{ k} K′ < r

1 P K− f dξ f dΞ δ − r ≥  ZS [0,K] ZS  × (r K)C (A.13) − + P f 2 dΞ >C ≤ rKδ2 r Z  −1 2 C R ′ f dξ (r K)C E[1 e− S×[0,K ] ] − + − −1 . ≤ rKδ2 1 e K′e−C − − POISSON REPRESENTATIONS 41

Proof. By (A.2) and the Chebyshev inequality, 2 1 f dΞ C 1 P K− f dξ f dΞ δ Ξ 1 + R f 2 dΞ>C , − ≥ ≤ Kδ2 ∧ ≤ Kδ2 { }  ZS [0,K] ZS  R × and taking expectations gives the first inequality in (A.12). The second inequality follows by (A.4). Similarly, for the second part, 2 1 (1 K/r)f dΞr P K− f dξ f dΞ δ Ξ − 1 − r ≥ r ≤ Kδ2 ∧  ZS [0,K] ZS  R × (r K)C 1 − + R f 2 dΞr>C , ≤ rKδ2 { } and taking expectations gives the first inequality in (A.13). The second inequality follows by (A.8). 

The estimates in Lemma A.11 allow verifying convergence of measure- valued processes satisfying (A.11) or (A.10) by verifying convergence of the corresponding particle representations.

n Theorem A.12. Let ξ be a sequence of cadlag hk (S [0, ))- { } M{ } × ∞ n valued processes satisfying (A.11) for cadlag hk (S)-valued processes Ξ . If the finite-dimensional distributions of ξn convergeM{ } to the finite-dimensional{ } distributions of ξ, then the finite-dimensional distributions of Ξn converge to the finite-dimensional distributions of Ξ satisfying R f dξ(t) Ξ K R (1 e−f ) dΞ(t) E[e− S×[0,K] ]= e− S − . |Ft n For n = 1, 2,..., let ξ be a cadlag hk (S [0, rn])-valued process sat- M{ } × n isfying (A.10) for cadlag hk (S)-valued processes Ξrn . If rn and the finite-dimensional distributionsM{ } of ξn converge to{ the finite-dimensional} →∞ n distributions of ξ, then the finite-dimensional distributions of Ξrn converge to the finite-dimensional distributions of Ξ satisfying R f dξ(t) Ξ K R (1 e−f ) dΞ(t) E[e− S×[0,K] ]= e− S − . |Ft Proof. Convergence of the finite-dimensional distributions follows eas- ily from the estimates in Lemma A.11. 

A.4. Martingale lemmas.

Lemma A.13. Let and be filtrations with . Suppose {Ft} {Gt} Gt ⊂ Ft that E[ X(t) + t Y (s) ds] < for each t, and | | 0 | | ∞ R t M(t)= X(t) Y (s) ds − Z0 42 T. G. KURTZ AND E. R. RODRIGUES is an -martingale. Then {Ft} t M(t)= E[X(t) ] E[Y (s) ] ds |Gt − |Gs Z0 is a t -martingale.c {G } Proof. Let D . Then ∈ Gt ⊂ Ft E[(M(t + r) M(t))1 ] − D t+r c= E E[Xc(t + r) ] E[X(t) ] E[Y (s) ] ds 1 |Gt+r − |Gt − |Gs D  Zt   t+r = E X(t + r) X(t) Y (s) ds 1 − − D  Zt   = 0, giving the martingale property. 

Lemma A.14. Let n be an increasing sequence of σ-algebras and X a sequence of random{F } variables satisfying E[sup X ] < and { n} n| n| ∞ limn Xn = X a.s. Then →∞

lim E[Xn n]= E X n . n |F F →∞  n  _ Proof. By the martingale convergence theorem, we have

E inf Xk n lim inf E[Xn n] lim supE[Xn n] E sup Xk n , k m F ≤ n |F ≤ n |F ≤ k m F  ≥ n  →∞ →∞  ≥ n  _ _ and the result follows by letting m .  →∞ A.5. Markov mapping theorem. The following theorem (extending Corol- lary 3.5 from [30]) plays an essential role in justifying the particle repre- sentations and can also be used to prove uniqueness for the correspond- ing measure-valued processes. Let (S, d) and (S0, d0) be complete, separable metric spaces, B(S) M(S) be the Banach space of bounded measurable ⊂ functions on S, with f = supx S f(x) and C(S) B(S) be the subspace of bounded continuousk functions.k ∈ An| operator| A ⊂B(S) B(S) is dissipa- ⊂ × tive if f1 f2 ε(g1 g2) f1 f2 for all (f1, g1), (f2, g2) A and ε> 0; A is a kpre-generator− − −if A isk ≥ dissipative k − k and there are sequences∈ of functions µ : S (S) and λ : S [0, ) such that for each (f, g) A n → P n → ∞ ∈

(A.14) g(x) = lim λn(x) (f(y) f(x))µn(x,dy) n − →∞ ZS POISSON REPRESENTATIONS 43 for each x S. A is graph separable if there exists a countable subset gk (A) C(S∈) such that the graph of A is contained in the bounded, pointwise{ }⊂ Dclosure∩ of the linear span of (g , Ag ) . [More precisely, we should say that { k k } there exists (gk, hk) A C(S) B(S) such that A is contained in the bounded pointwise{ closure}⊂ of∩ (g ,× h ) , but typically A is single-valued, so { k k } we use the more intuitive notation Agk.] These two conditions are satisfied by essentially all operators A that might reasonably be thought to be generators of Markov processes. Note that A is graph separable if A L L, where L B(S) is separable in the sup norm topology, for example,⊂ if ×S is locally compact,⊂ and L is the space of continuous functions vanishing at infinity. A collection of functions D C(S) is separating if ν,µ (S) and f dν = ⊂ ∈ P S S f dµ for all f D imply µ = ν. ∈ Y R R For an S0-valued, measurable process Y , t will denote the completion of the σ-algebra σ(Y (0), r h(Y (s)) ds, r t,F h B(S )). For almost every 0 ≤ ∈ 0 Y b Y Y t, Y (t) will be t -measurable,R but in general, t does not contain t = F TY Y F F σ(Y (s) : s t). Let = t : Y (t) is t measurable . If Y is cadlag and has no fixed points≤ b of discontinuity{ [i.e.,F for every tb, Y}(t)= Y (t ) a.s.], then Y − T = [0, ). DS[0, ) denotes the spaceb of cadlag, S-valued functions with ∞ ∞ the Skorohod topology, and MS[0, ) denotes the space of Borel measurable functions, x : [0, ) S, topologized∞ by convergence in Lebesgue measure. ∞ →

Theorem A.15. Let (S, d) and (S0, d0) be complete, separable metric spaces. Let A C(S) C(S) and ψ C(S), ψ 1. Suppose that for each f (A) there⊂ exists c× > 0 such that∈ ≥ ∈ D f (A.15) Af(x) c ψ(x), x A, | |≤ f ∈ and define A0f(x)= Af(x)/ψ(x). Suppose that A0 is a graph-separable pre-generator, and suppose that (A)= (A ) is closed under multiplication and is separating. Let γ : S D D 0 → S0 be Borel measurable, and let α be a transition function from S0 into S [y S0 α(y, ) (S) is Borel measurable] satisfying h γ(z)α(y, dz)= ∈ → · ∈ P 1 ◦ h(y), y S0, h B(S0), that is, α(y, γ− (y)) = 1. AssumeR that ψ(y) ψ(z)α∈(y, dz) <∈ for each y S , and define ≡ S ∞ ∈ 0 e R C = f(z)α( , dz), Af(z)α( , dz) : f (A) . · · ∈ D ZS ZS   Let µ (S ), and define ν = α(y, )µ (dy). 0 ∈ P 0 0 · 0 (a) If Y satisfies t E[ψ(Y (s))] dsR < for all t 0, and Y is a solution of 0 ∞ ≥ the martingale problem for (C,µ0), then there exists a solution X of the R martingalee probleme fore (A, ν0) such that Y has the samee distribution on

MS0 [0, ) as Y = γ X. If Y and Y are cadlag, then Y and Y have the same distribution∞ on◦D [0, ). e S0 ∞ e e 44 T. G. KURTZ AND E. R. RODRIGUES

(b) For t TY , ∈ (A.16) P X(t) Γ Y = α(Y (t), Γ), Γ (S). { ∈ |Ft } ∈B (c) If, in addition, uniqueness holds for the martingale problem for (A, ν ), b 0 then uniqueness holds for the M [0, )-martingale problem for (C,µ ). S0 ∞ 0 If Y has sample paths in DS0 [0, ), then uniqueness holds for the D [0, )-martingale problem for (∞C,µ ). S0 ∞ 0 (d) If uniquenesse holds for the martingale problem for (A, ν0), then Y re- stricted to TY is a Markov process.

Remark A.16. Theorem A.15 can be extended to cover a large class of generators whose range contains discontinuous functions. (See [30], Corollary 3.5 and Theorem 2.7.) In particular, suppose A1,...,Am satisfy the condi- tions of Theorem A.15 for a common domain = (A )= = (A ), D D 1 · · · D m and β1,...,βm are nonnegative functions in M(S). Then the conclusions of Theorem A.15 hold for Af = β A f + + β A f. 1 1 · · · m m Proof of Theorem A.15. Theorem 3.2 of [30] can be extended to operators satisfying (A.15) by applying Corollary 1.12 of [34] (with the op- erator B in that corollary set equal zero) in place of Theorem 2.6 of [30]. Alternatively, see Corollary 3.3 of [32]. 

A.6. Uniqueness for martingale problems. Assume that B C(E) C(E), that (B) is closed under multiplication and is separating,⊂ and× D that existence and uniqueness hold for the DE[0, ) martingale problem for (B,ν) for each initial distribution ν (E). Without∞ loss of generality, we can assume g (B) satisfies 0 g ∈1. P By Theorem 4.10.1∈ D of [9], existence≤ and≤ uniqueness then follows for the n-particle motion martingale problem with generator n n n Bg(xi) (A.17) B = f(x),f(x) : f(x)= g(xi) . ( g(xi) ! ) Xi=1 Yi=1 Actually, the cited theorem implies uniqueness for the ordered n-particle motion with generator n n n Bgi(xi) B = f(x),f(x) : f(x)= gi(xi), gi (B) , ( gi(xi) ! ∈ D ) Xi=1 Yi=1 but Theoreme A.15 can be applied to obtain uniqueness for Bn from unique- n n ness for B . Define γ(x)= i=1 δxi and let α(y, ) average over all permu- n · tations of the xi in y = δx . i=1P i e P POISSON REPRESENTATIONS 45

Now consider a generator for a process with state space n S = δ : x E , xi i ∈ n ( ) [ Xi=1 where we allow n = 0, that is, no particles exist. n For f(x,n)= i=1 g(xi), let Af(x,n)= BnQf(x,n) k + f(x,n) λk(x) g(zi) 1 ηk(x, dz1,...,dzk) Ek − ! Xk Z Yi=1 1 + f(x,n) µ(x, z1,...,zl) l 1 , g(zi) − (z1,...,zl) x1,...,xn  i=1  X⊂{ } Q k where λk,µ 0 and ηk is a transition function from S to E . The generator has the following≥ simple interpretation: in between birth and death events the particles move independently with motion determined by B. At rate λk(x), k new particles are created with locations in E determined by ηk. At rate µ(x, z1,...,zl), the particles at z1,...,zl are removed. Let

β(x)= λk(x)+ µ(x, z1,...,zl).

k (z1,...,z ) x1,...,xn X l X⊂{ } Then for each initial distribution ν0 and each m> 0, a localization argument and Theorem 4.10.3 of [9] imply existence and uniqueness of the martingale problem for (A, ν0) up to the first time the solution leaves x : β(x) 0. Essentially the≤ same argument∞ gives existence and uniqueness for gener- ators of the form (3.1) provided inf (a(x)r b(x)) > 0 and there exists a x − solution satisfying sups t a(Xi(s)) < a.s. for each t> 0. ≤ ∞ P REFERENCES [1] Berestycki, J., Berestycki, N. and Limic, V. The λ-coalescent speed of coming down from infinity. Preprint. [2] Bertoin, J. and Le Gall, J.-F. (2006). Stochastic flows associated to coalescent pro- cesses. III. Limit theorems. Illinois J. Math. 50 147–181 (electronic). MR2247827 [3] Bhattacharya, R. N. (1982). On the functional and the law of the iterated logarithm for Markov processes. Z. Wahrsch. Verw. Gebiete 60 185–201. MR0663900 [4] Dawson, D. A. (1975). Stochastic evolution equations and related measure processes. J. Multivariate Anal. 5 1–52. MR0388539 46 T. G. KURTZ AND E. R. RODRIGUES

[5] Donnelly, P. and Kurtz, T. G. (1996). A countable representation of the Fleming– Viot measure-valued diffusion. Ann. Probab. 24 698–742. MR1404525 [6] Donnelly, P. and Kurtz, T. G. (1999). Genealogical processes for Fleming–Viot models with selection and recombination. Ann. Appl. Probab. 9 1091–1148. MR1728556 [7] Donnelly, P. and Kurtz, T. G. (1999). Particle representations for measure-valued population models. Ann. Probab. 27 166–205. MR1681126 [8] Ethier, S. N. and Griffiths, R. C. (1993). The transition function of a measure- valued branching diffusion with immigration. In Stochastic Processes 71–79. Springer, New York. MR1427302 [9] Ethier, S. N. and Kurtz, T. G. (1986). Markov Processes: Characterization and Convergence. Wiley, New York. MR0838085 [10] Evans, S. N. (1993). Two representations of a conditioned superprocess. Proc. Roy. Soc. Edinburgh Sect. A 123 959–971. MR1249698 [11] Evans, S. N. and O’Connell, N. (1994). Weighted occupation time for branching particle systems and a representation for the supercritical superprocess. Canad. Math. Bull. 37 187–196. MR1275703 [12] Evans, S. N. and Perkins, E. (1990). Measure-valued Markov branching processes conditioned on nonextinction. Israel J. Math. 71 329–337. MR1088825 [13] Fitzsimmons, P. J. (1988). Construction and regularity of measure-valued Markov branching processes. Israel J. Math. 64 337–361 (1989). MR0995575 [14] Gorostiza, L. G. and Lopez-Mimbela,´ J. A. (1990). The multitype measure branching process. Adv. in Appl. Probab. 22 49–67. MR1039376 [15] Grey, D. R. (1988). Supercritical branching processes with density independent catastrophes. Math. Proc. Cambridge Philos. Soc. 104 413–416. MR0948925 [16] Grimvall, A. (1974). On the convergence of sequences of branching processes. Ann. Probab. 2 1027–1045. MR0362529 [17] Harris, T. E. (1951). Some mathematical models for branching processes. In Pro- ceedings of the Second Berkeley Symposium on and Prob- ability, 1950 305–328. Univ. of California Press, Berkeley and Los Angeles. MR0045331 [18] Helland, I. S. (1981). Minimal conditions for weak convergence to a diffusion process on the line. Ann. Probab. 9 429–452. MR0614628 [19] Hering, H. (1978). The non-degenerate limit for supercritical branching diffusions. Duke Math. J. 45 561–600. MR0507459 [20] Hering, H. and Hoppe, F. M. (1981). Critical branching diffusions: Proper normal- ization and conditioned limit. Ann. Inst. H. Poincar´eSect. B (N.S.) 17 251–274. MR0631242 [21] Ikeda, N., Nagasawa, M. and Watanabe, S. (1965). On branching Markov pro- cesses. Proc. Japan Acad. 41 816–821. MR0202195 [22] Ikeda, N., Nagasawa, M. and Watanabe, S. (1968). Branching Markov processes. I. J. Math. Kyoto Univ. 8 233–278. MR0232439 [23] Ikeda, N., Nagasawa, M. and Watanabe, S. (1968). Branching Markov processes. II. J. Math. Kyoto Univ. 8 365–410. MR0238401 [24] Ikeda, N., Nagasawa, M. and Watanabe, S. (1969). Branching Markov processes. III. J. Math. Kyoto Univ. 9 95–160. MR0246376 [25] Joffe, A. and Metivier,´ M. (1986). Weak convergence of sequences of semimartin- gales with applications to multitype branching processes. Adv. in Appl. Probab. 18 20–65. MR0827331 POISSON REPRESENTATIONS 47

[26] Keiding, N. (1975). Extinction and exponential growth in random environments. Theoret. Population Biol. 8 49–63. [27] Kulperger, R. (1979). Brillinger type conditions for a simple branching diffusion process. Stochastic Process. Appl. 9 55–66. MR0544715 [28] Kurtz, T. G. (1973). A limit theorem for perturbed operator semigroups with ap- plications to random evolutions. J. Funct. Anal. 12 55–67. MR0365224 [29] Kurtz, T. G. (1978). Diffusion approximations for branching processes. In Branching Processes (Conf., Saint Hippolyte, Que., 1976). Adv. Probab. Related Topics 5 269–292. Dekker, New York. MR0517538 [30] Kurtz, T. G. (1998). Martingale problems for conditional distributions of Markov processes. Electron. J. Probab. 3 29 pp. (electronic). MR1637085 [31] Kurtz, T. G. (2000). Particle representations for measure-valued population pro- cesses with spatially varying birth rates. In Stochastic Models (Ottawa, ON, 1998). CMS Conf. Proc. 26 299–317. Amer. Math. Soc., Providence, RI. MR1765017 [32] Kurtz, T. G. and Nappo, G. (2010). The filtered martingale problem. In Handbook on Nonlinear Filtering (D. Crisan and B. Rozovsky, eds.). Oxford Univ. Press. To appear. [33] Kurtz, T. G. and Protter, P. (1991). Weak limit theorems for stochastic integrals and stochastic differential equations. Ann. Probab. 19 1035–1070. MR1112406 [34] Kurtz, T. G. and Stockbridge, R. H. (2001). Stationary solutions and forward equations for controlled and singular martingale problems. Electron. J. Probab. 6 52 pp. (electronic). MR1873294 [35] Kurtz, T. G. and Xiong, J. (1999). Particle representations for a class of nonlinear SPDEs. Stochastic Process. Appl. 83 103–126. MR1705602 [36] Lamperti, J. and Ney, P. (1968). Conditioned branching processes and their limit- ing diffusions. Teor. Verojatnost. i Primenen. 13 126–137. MR0228073 [37] Li, Z. H. (1992). Measure-valued branching processes with immigration. Stochastic Process. Appl. 43 249–264. MR1191150 [38] Li, Z. H., Li, Z. B. and Wang, Z. K. (1993). Asymptotic behavior of the measure- valued branching process with immigration. Sci. China Ser. A 36 769–777. MR1247000 [39] Li, Z. and Wang, Z. (1999). Measure-valued branching processes and immigration processes. Adv. Math. (China) 28 105–134. MR1723025 [40] Mellein, B. (1982). Diffusion limits of conditioned critical Galton–Watson processes. Rev. Colombiana Mat. 16 125–140. MR0685248 [41] Pakes, A. G. (1986). The Markov branching-catastrophe process. Stochastic Process. Appl. 23 1–33. MR0866285 [42] Pakes, A. G. (1987). Limit theorems for the population size of a birth and death process allowing catastrophes. J. Math. Biol. 25 307–325. MR0900324 [43] Pakes, A. G. (1988). The Markov branching process with density-independent catas- trophes. I. Behaviour of extinction . Math. Proc. Cambridge Philos. Soc. 103 351–366. MR0923688 [44] Pakes, A. G. (1989). Asymptotic results for the extinction time of Markov branch- ing processes allowing emigration. I. decrements. Adv. in Appl. Probab. 21 243–269. MR0997723 [45] Pakes, A. G. (1989). The Markov branching process with density-independent catas- trophes. II. The subcritical and critical cases. Math. Proc. Cambridge Philos. Soc. 106 369–383. MR1002548 48 T. G. KURTZ AND E. R. RODRIGUES

[46] Pakes, A. G. (1990). The Markov branching process with density-independent catas- trophes. III. The supercritical case. Math. Proc. Cambridge Philos. Soc. 107 177–192. MR1021881 [47] Schweinsberg, J. (2000). A necessary and sufficient condition for the Λ-coalescent to come down from infinity. Electron. Comm. Probab. 5 1–11 (electronic). MR1736720 [48] Stannat, W. (2003). On transition semigroups of (A, Ψ)-superprocesses with immi- gration. Ann. Probab. 31 1377–1412. MR1989437 [49] Watanabe, S. (1968). A limit theorem of branching processes and continuous state branching processes. J. Math. Kyoto Univ. 8 141–167. MR0237008

Departments of Mathematics Instituto de Matematicas´ and Statistics UNAM University of Wisconsin, Madison Mexico,´ DF 04510 480 Lincoln Drive Mexico Madison, Wisconsin 53706-1388 E-mail: [email protected] USA [email protected] E-mail: [email protected]