This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Anomalous spreading in reducible multitype
branching Brownian motion

Mohamed Ali Belloum belloum@math.univ-paris13.fr, Université Sorbonne Paris Nord, LAGA, CNRS, UMR 7539, F-93430, Villetaneuse, France.    Bastien Mallein mallein@math.univ-paris13.fr, Université Sorbonne Paris Nord, LAGA, UMR 7539, F-93430, Villetaneuse, France.
Abstract

We consider a two-type reducible branching Brownian motion, defined as a particle system on the real line in which particles of two types move according to independent Brownian motion and create offspring at constant rate. Particles of type 11 can give birth to particles of types 11 and 22, but particles of type 22 only give birth to descendants of type 22. Under some specific conditions, Biggins [Big12] shows that this process exhibit an anomalous spreading behaviour: the rightmost particle at time tt is much further than the expected position for the rightmost particle in a branching Brownian motion consisting only of particles of type 11 or of type 22. This anomalous spreading also has been investigated from a reaction-diffusion equation standpoint by Holzer [Hol14, Hol16]. The aim of this article is to study the asymptotic behaviour of the position of the furthest particles in the two-type reducible branching Brownian motion, obtaining in particular tight estimates for the median of the maximal displacement.

1 Introduction

The standard branching Brownian motion is a particle system on the real line that can be constructed as follows. It starts with a unique particle at time 0, that moves according to a standard Brownian motion. After an exponential time of parameter 11, this particle dies and is replaced by two children. The two daughter particles then start independent copies of the branching Brownian motion from their current position. For all t0t\geq 0, we write 𝒩t\mathcal{N}_{t} the set of particles alive at time tt, and for u𝒩tu\in\mathcal{N}_{t} we denote by Xu(t)X_{u}(t) the position at time tt of that particle.

The branching Brownian motion (or BBM) is strongly related to the F-KPP reaction-diffusion equation, defined as

tu=12Δuu(1u).\partial_{t}u=\frac{1}{2}\Delta u-u(1-u). (1.1)

More precisely, given a measurable function f:[0,1]f:\mathbb{R}\to[0,1], set for xx\in\mathbb{R} and t0t\geq 0

ut(x)=𝐄(u𝒩tf(Xu(t)+x)),u_{t}(x)=\mathbf{E}\left(\prod_{u\in\mathcal{N}_{t}}f(X_{u}(t)+x)\right),

then uu is the solution of (1.1) with u0(x)=f(x)u_{0}(x)=f(x). In particular, setting Mt=maxu𝒩tXu(t)M_{t}=\max_{u\in\mathcal{N}_{t}}X_{u}(t), we note that the tail distribution of Mt-M_{t} is the solution at time tt of (1.1) with u0(z)=𝟙{z<0}u_{0}(z)=\mathbbm{1}_{\left\{z<0\right\}}.

Thanks to this observation, Bramson [Bra78] obtained an explicit formula for the asymptotic behaviour of the median of MtM_{t}. Precisely, he observed that setting

mt:=2t322logtm_{t}:=\sqrt{2}t-\frac{3}{2\sqrt{2}}\log t (1.2)

the process (Mtmt,t0)(M_{t}-m_{t},t\geq 0) is tight. Lalley and Sellke [LS87] refined this result and prove that MtmtM_{t}-m_{t} converges in law toward a Gumbel random variable shifted by an independent copy of 12logZ\frac{1}{\sqrt{2}}\log Z_{\infty}, where ZZ_{\infty} is the a.s. limit as tt\to\infty of the derivative martingale, defined by

Zt:=u𝒩t(2tXu(t))e2Xu(t)2t.Z_{t}:=\sum_{u\in\mathcal{N}_{t}}(\sqrt{2}t-X_{u}(t))e^{\sqrt{2}X_{u}(t)-2t}.

The derivative martingale is called that way due to its relationship to the derivative at its critical point of the additive martingale introduced by McKean in [McK75], defined as

Wt(θ):=u𝒩teθXu(t)t(1+θ22).W_{t}(\theta):=\sum_{u\in\mathcal{N}_{t}}e^{\theta X_{u}(t)-t(1+\frac{\theta^{2}}{2})}.

It was shown in [Nev88] that (Wt(θ),t0)(W_{t}(\theta),t\geq 0) is uniformly integrable if and only if |θ|<2|\theta|<\sqrt{2} and converges to an a.s. positive limit W(θ)W_{\infty}(\theta) in that case. Otherwise, it converges to 0 a.s. This result has later been extended by Biggins [Big77] and Lyons [Lyo97] to the branching random walk, which is a discrete-time analogous to the BBM.

The behaviour of the particles at the tip of branching Brownian motions was later investigated by Aidékon, Berestycki, Brunet and Shi [ABBS13] as well as Arguin, Bovier and Kistler [ABK11, ABK12, ABK13]. They proved that the centred extremal process of the standard BBM, defined by

^tR=u𝒩tδXu(t)mt\widehat{\mathcal{E}}^{R}_{t}=\sum_{u\in\mathcal{N}_{t}}\delta_{X_{u}(t)-m_{t}}

converges in distribution to a decorated Poisson point process with (random) intensity 2cZe2xdx\sqrt{2}c_{\star}Z_{\infty}e^{-\sqrt{2}x}\mathrm{d}x. More precisely, there exists a law 𝔇\mathfrak{D} on point measures such that writing (Dj,j1)(D_{j},j\geq 1) i.i.d. point measures with law 𝔇\mathfrak{D} and (ξj,j0)(\xi_{j},j\geq 0) the atoms of an independent Poisson point process with intensity 2ce2xdx\sqrt{2}c_{\star}e^{-\sqrt{2}x}\mathrm{d}x, which are further independent of ZZ_{\infty}, and defining

=j1dDjδξj+d+12logZ,\mathcal{E}_{\infty}=\sum_{j\geq 1}\sum_{d\in D_{j}}\delta_{\xi_{j}+d+\frac{1}{\sqrt{2}}\log Z_{\infty}},

we have limtt=\lim_{t\to\infty}\mathcal{E}_{t}=\mathcal{E}_{\infty} in law, for the topology of the vague convergence. We give more details on these results in Section 3.

We refer to the above limit as a decorated Poisson point process, or DPPP(2cZe2xdx,𝔇)\text{DPPP}(\sqrt{2}c_{\star}Z_{\infty}e^{-\sqrt{2}x}\mathrm{d}x,\mathfrak{D}). Maillard [Mai13] obtained a characterization of this type of point processes as satisfying a stability by superposition property. This characterization was used in [Mad17] to prove a similar convergence in distribution to a DPPP for the shifted extremal process of the branching random walk. Subag and Zeitouni [SZ15] studied in more details the family of shifted randomly decorated Poisson random measures with exponential intensity.

In this article, we take interest in the two-type reducible branching Brownian motion. This is a particle system on the real line in which particles possess a type in addition with their position. Particles of type 11 move according to Brownian motions with diffusion coefficient σ12\sigma^{2}_{1} and branch at rate β1\beta_{1} into two children of type 11. Additionally, they give birth to particles of type 22 at rate α\alpha. Particles of type 22 move according to Brownian motions with diffusion coefficient σ22\sigma^{2}_{2} and branch at rate β2\beta_{2}, but cannot give birth to descendants of type 11.

In [Big12], Biggins observe that in some cases multitype reducible branching random walks exhibit an anomalous spreading property. Precisely, the rightmost particle at time tt is shown to be around position vtvt, with the speed vv of the two-type process being larger than the speed of a branching random walk consisting only of particles of type 11 or uniquely of particles of type 22. Therefore, the multitype system can invade its environment at a higher speed than the one that either particles of type 11 or particles of type 22 would be able to sustain on their own.

Holzer [Hol14, Hol16] extended the results of Biggins to this setting, by considering the associated system of F-KPP equations, describing the speed of the rightmost particle in the system in terms of σ1,β1,σ2\sigma_{1},\beta_{1},\sigma_{2} and β2\beta_{2} (the parameter α\alpha does not modify the speed of the two-type particle system). Our aim is to study in more details the position of the maximal displacement, in particular in the case when anomalous spreading occurs, for this two type BBM. We also take interest in the extremal process formed by the particles of type 22 at time tt, and show it to converge towards a DPPP.

Recall that the reducible two-type BBM is defined by five parameters, the diffusion coefficient σ12\sigma_{1}^{2}, σ22\sigma_{2}^{2} of particles of type 11 and 22, their branching rate β1\beta_{1}, β2\beta_{2}, and the rate α\alpha at which particles of type 11 create particles of type 22. However, up to a dilation of time and space, it is possible to modify these parameters in such a way that σ22=β2=1\sigma_{2}^{2}=\beta_{2}=1. Additionally, the parameter α\alpha plays no role in the value of the speed of the multitype process. We can therefore describe the phase space of this process in terms of the two parameters σ2:=σ12\sigma^{2}:=\sigma_{1}^{2} and β=β1\beta=\beta_{1}, and identify for which parameters does anomalous spreading occurs. This is done in Figure 1.

β\betaσ2\sigma^{2}11𝒞II\mathcal{C}_{II}𝒞I\mathcal{C}_{I}𝒞III\mathcal{C}_{III}
Figure 1: Phase diagram of the two-type reducible BBM.

We decompose the state space (β,σ2)+2(\beta,\sigma^{2})\in\mathbb{R}_{+}^{2} into three regions:

𝒞I\displaystyle\mathcal{C}_{I} ={(β,σ2):σ2>𝟙{β1}β+𝟙{β>1}β2β1}\displaystyle=\left\{(\beta,\sigma^{2}):\sigma^{2}>\frac{\mathbbm{1}_{\left\{\beta\leq 1\right\}}}{\beta}+\mathbbm{1}_{\left\{\beta>1\right\}}\frac{\beta}{2\beta-1}\right\}
𝒞II\displaystyle\mathcal{C}_{II} ={(β,σ2):σ2<𝟙{β1}β+𝟙{β>1}(2β)}\displaystyle=\left\{(\beta,\sigma^{2}):\sigma^{2}<\frac{\mathbbm{1}_{\left\{\beta\leq 1\right\}}}{\beta}+\mathbbm{1}_{\left\{\beta>1\right\}}(2-\beta)\right\}
𝒞III\displaystyle\mathcal{C}_{III} ={(β,σ2):σ2+β>2 and σ2<β2β1}.\displaystyle=\left\{(\beta,\sigma^{2}):\sigma^{2}+\beta>2\text{ and }\sigma^{2}<\frac{\beta}{2\beta-1}\right\}.

If (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, the speed of the two-type reducible BBM is 2βσ2\sqrt{2\beta\sigma^{2}}, which is the same as particles of type 11 alone, ignoring births of particles of type 22. Thus in this situation, the asymptotic behaviour of the extremal process is dominated by the long-time behaviour of particles of type 11. Conversely, if (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, then the speed of the process is 2\sqrt{2}, equal to the speed of a single BBM of particles of type 22. In that situation, the asymptotic behaviour of particles of type 22 dominates the extremal process. Finally, if (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III}, the speed of the process is larger than max(2,2βσ2)\max(\sqrt{2},\sqrt{2\beta\sigma^{2}}), and we will show that in this case the extremal process will be given by a mixture of the long-time asymptotic of the processes of particles of type 11 and 22.

For all t0t\geq 0, we write 𝒩t\mathcal{N}_{t} the set of all particles alive at time tt, as well as 𝒩t1\mathcal{N}^{1}_{t} and 𝒩t2\mathcal{N}^{2}_{t} the set of particles of type 11 and type 22 respectively. We also write Xu(t)X_{u}(t) for the position at time tt of u𝒩tu\in\mathcal{N}_{t}, and for all sts\leq t, Xu(s)X_{u}(s) for the position of the ancestor at time ss of particle uu. If u𝒩t2u\in\mathcal{N}^{2}_{t}, we denote by T(u)T(u) the time at which the oldest ancestor of type 22 of uu was born. In this article, we study the asymptotic behaviour of the extremal process of particles of type 22 in this 22-type BBM for Lebesgue-almost every values of σ2\sigma^{2} and β\beta (and for α(0,)\alpha\in(0,\infty)).

We divide the main result of our article into three theorems, one for each area the pair (β,σ2)(\beta,\sigma^{2}) belongs to. We begin with the asymptotic behaviour of extremal particles when (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, in which case the extremal point measure is similar to the one observed in a branching Brownian motion of particles of type 11.

Theorem 1.1 (Domination of particles of type 11).

If (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, then there exist a constant c(I)>0c_{(I)}>0 and a point measure distribution 𝔇(I)\mathfrak{D}^{(I)} such that setting mt(I):=2σ2βt322β/σ2logtm^{(I)}_{t}:=\sqrt{2\sigma^{2}\beta}t-\frac{3}{2\sqrt{2\beta/\sigma^{2}}}\log t we have

limtu𝒩t2δXu(t)mt(I)=(I) in law,\lim_{t\to\infty}\sum_{u\in\mathcal{N}_{t}^{2}}\delta_{X_{u}(t)-m^{(I)}_{t}}=\mathcal{E}^{(I)}_{\infty}\quad\text{ in law},

for the topology of the vague convergence, where (I)\mathcal{E}^{(I)}_{\infty} is a DPPP(2β/σ2c(I)Z(1)e2β/σ2xdx,𝔇(I)\sqrt{2\beta/\sigma^{2}}c_{(I)}Z^{(1)}_{\infty}e^{-\sqrt{2\beta/\sigma^{2}}x}\mathrm{d}x,\mathfrak{D}^{(I)}), where

Z(1):=limtu𝒩t1(2σ2βtXu(t))e2β/σ2Xu(t)2βta.s.Z^{(1)}_{\infty}:=\lim_{t\to\infty}\sum_{u\in\mathcal{N}^{1}_{t}}(\sqrt{2\sigma^{2}\beta}t-X_{u}(t))e^{\sqrt{2\beta/\sigma^{2}}X_{u}(t)-2\beta t}\quad\text{a.s.}

Additionally, we have limt𝐏(Mtmt(I)+x)=𝐄(ec(I)Z(1)e2β/σ2x)\displaystyle\lim_{t\to\infty}\mathbf{P}(M_{t}\leq m^{(I)}_{t}+x)=\mathbf{E}\left(e^{-c_{(I)}Z^{(1)}_{\infty}e^{-\sqrt{2\beta/\sigma^{2}}x}}\right) for all xx\in\mathbb{R}.

We underscore that in this theorem, the values of c(I)c_{(I)} and 𝔇(I)\mathfrak{D}^{(I)} are obtained implicitly, and depend on the parameters α\alpha, β\beta and σ2\sigma^{2} of the multitype branching Brownian motion. The identification of the law of the extremal point measure (I)\mathcal{E}^{(I)} is based on the fact that it satisfies a stability by superposition property, and using Maillard’s characterization [Mai13].

If (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, we show that the extremal process of the two-type BBM is similar to the extremal process of a single BBM of particles of type 22, up to a random shift whose law depend on the behaviour of particles of type 11.

Theorem 1.2 (Domination of particles of type 22).

If (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, then writing c>0c_{\star}>0 the prefactor of the intensity measure in the extremal process of the standard BBM and 𝔇\mathfrak{D} the law of its decoration, setting mt(II):=mt=2t322logtm^{(II)}_{t}:=m_{t}=\sqrt{2}t-\frac{3}{2\sqrt{2}}\log t, we have

limtu𝒩t2δXu(t)mt(II)=(II) in law,\lim_{t\to\infty}\sum_{u\in\mathcal{N}_{t}^{2}}\delta_{X_{u}(t)-m^{(II)}_{t}}=\mathcal{E}^{(II)}_{\infty}\quad\text{ in law},

for the topology of the vague convergence, where (II)\mathcal{E}^{(II)}_{\infty} is a DPPP(2cZ¯e2xdx,𝔇\sqrt{2}c_{\star}\overline{Z}_{\infty}e^{-\sqrt{2}x}\mathrm{d}x,\mathfrak{D}) and Z¯\overline{Z}_{\infty} is defined in Lemma 5.3. Additionally, for all xx\in\mathbb{R} we have limt𝐏(Mtmt(II)+x)=𝐄(ecZ¯e2x).\displaystyle\lim_{t\to\infty}\mathbf{P}(M_{t}\leq m^{(II)}_{t}+x)=\mathbf{E}\left(e^{-c_{\star}\overline{Z}_{\infty}e^{-\sqrt{2}x}}\right).

We finally take interest in the case (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III}, in this situation the BBM exhibits an anomalous spreading behaviour. The extremal process contains only particles of type 22, but particles travel at greater speed that would have been observed in a BBM of particles of type 11 or of type 22.

Theorem 1.3 (Anomalous spreading).

If (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III}, then setting

mt(III)=σ2β2(1σ2)(β1)tandθ=2β11σ2,m^{(III)}_{t}=\frac{\sigma^{2}-\beta}{\sqrt{2(1-\sigma^{2})(\beta-1)}}t\quad\text{and}\quad\theta=\sqrt{2\frac{\beta-1}{1-\sigma^{2}}},

we have

limtu𝒩t2δXu(t)mt(III)=(III) in law,\lim_{t\to\infty}\sum_{u\in\mathcal{N}_{t}^{2}}\delta_{X_{u}(t)-m^{(III)}_{t}}=\mathcal{E}^{(III)}_{\infty}\quad\text{ in law},

for the topology of the vague convergence, where \mathcal{E}_{\infty} is a DPPP(θc(III)W(θ)eθxdx,𝔇(III)\theta c_{(III)}W_{\infty}(\theta)e^{-\theta x}\mathrm{d}x,\mathfrak{D}^{(III)}) and

  • W(θ)=limtu𝒩t1eθXu(t)t(β+θ2σ2/2)W_{\infty}(\theta)=\lim_{t\to\infty}\sum_{u\in\mathcal{N}^{1}_{t}}e^{\theta X_{u}(t)-t(\beta+\theta^{2}\sigma^{2}/2)} is the a.s. limit of an additive martingale of the BBM of particles of type 11 with parameter θ\theta,

  • c(III)=αC(θ)2(β1)c_{(III)}=\frac{\alpha C(\theta)}{2(\beta-1)} with the function CC being defined in (3.16),

  • 𝔇(III)\mathfrak{D}^{(III)} is the law of the point measure 𝒟θ\mathcal{D}^{\theta} defined in (3.19).

Additionally, for all xx\in\mathbb{R} we have limt𝐏(Mtmt(III)+x)=𝐄(ec(III)W(θ)eθx).\displaystyle\lim_{t\to\infty}\mathbf{P}(M_{t}\leq m^{(III)}_{t}+x)=\mathbf{E}\left(e^{-c_{(III)}W_{\infty}(\theta)e^{-\theta x}}\right).

Remark 1.4.

Contrarily to what happens in Theorems 1.1 and 1.2, the extremal process obtained in Theorem 1.3 is not shifted by a random variable associated to a derivative martingale, but by an additive martingale of the BBM. Additionally, it is worth noting that contrarily to the median of the maximal displacements in domains 𝒞I\mathcal{C}_{I} and 𝒞II\mathcal{C}_{II}, when anomalous spreading occurs there is no logarithmic correction in the median of the maximal displacement.

Remark 1.5.

Observe that in Theorems 1.11.3, we obtain the convergence in law for the topology of the vague convergence of extremal processes to DPPPs as well as the convergence in law of their maximum to the maximum of this DPPP. These two convergences can be synthesized into the joint convergence of the extremal process together with its maximum, which is equivalent to the convergence of t,φ{\left\langle\mathcal{E}_{t},\varphi\right\rangle} for all continuous bounded functions φ\varphi with bounded support on the left (see e.g. [BBCM20, Lemma 4.4]).

The rest of the article is organized as follows. We discuss our results in the next section, by putting them in the context of the state of the art for single type and multitype branching processes, and for coupled reaction-diffusion equations. In Section 3, we introduce part of the notation and results on branching Brownian motions that will be needed in our proofs, in particular for the definition of the decorations laws of the extremal process. We introduce in Section 4 a multitype version of the celebrated many-to-one lemma. Finally, we prove Theorem 1.2 in Section 5, Theorem 1.1 in Section 6 and Theorem 1.3 in Section 7.

2 Discussion of our main result

We compare our main results for the asymptotic behaviour of the two-type reducible branching Brownian motion to the pre-existing literature. We begin by introducing the optimization problem associated to the computation of the speed of the rightmost particle in this process. Loosely speaking, this optimization problem is related to the “choice” of the time between 0 and tt and position at which the ancestral lineage leading to one of the rightmost positions switches from type 11 to type 22. The optimization problem was introduced by Biggins [Big12] for the computation of the speed of multitype reducible branching random walks. It allows us to describe the heuristics behind the main theorems.

We then compare Theorem 1.3 to the results obtained on the extremal process of time-inhomogeneous branching Brownian motions, and in particular with the results of Bovier and Hartung [BH14]. In Section 2.3, we apply our results to the work of Holzer [Hol14, Hol16] on coupled F-KPP equations. We end this section with the discussion of further questions of interest for multitype reducible BBMs and some conjectures and open questions.

2.1 Associated optimization problem and heuristic

Despite the fact that spatial multitype branching processes have a long history, the study of the asymptotic behaviour of their largest displacement has not been considered until recently. As previously mentioned, Biggins [Big12] computed the speed of multitype reducible branching random walks. This process is a discrete-time particles system in which each particle gives birth to offspring in an independent fashion around its position, with a reproduction law that depends on its type, under the assumption that the Markov chain associated to the type of a typical individual in the process is reducible. Ren and Yang [RY14] then considered the asymptotic behaviour of the maximal displacement in an irreducible multitype BBM.

In [Big10], Biggins gives an explicit description of the speed of a reducible two-type branching random walk as the solution of an optimization problem. In the context of the two-type BBM we consider, the optimization problem can be expressed as such:

v=max{pa+(1p)b:p[0,1],p(a22σ2β)0,p(a22σ2β)+(1p)(b221)0}.v=\max\left\{pa+(1-p)b:p\in[0,1],\ p\left(\frac{a^{2}}{2\sigma^{2}}-\beta\right)\leq 0,\ p\left(\frac{a^{2}}{2\sigma^{2}}-\beta\right)+(1-p)\left(\frac{b^{2}}{2}-1\right)\leq 0\right\}. (2.1)

This optimization problem can be understood as follows. It is well-known that if a<2σ2βa<\sqrt{2\sigma^{2}\beta} and b2b\geq\sqrt{2}, there are with high probability around ept(βa2/2σ2)+o(t)e^{pt(\beta-a^{2}/2\sigma^{2})+o(t)} particles of type 11 at time ptpt to the right of position ptapta, and a typical particle of type 22 has probability e(1p)t(1b2/2)+o(t)e^{(1-p)t(1-b^{2}/2)+o(t)} of having a descendant to the right of position (1p)bt(1-p)bt at time (1p)t(1-p)t. Therefore, for all (p,a,b)(p,a,b) such that

p[0,1],p(a22σ2β)0,p(a22σ2β)+(1p)(b221)0,p\in[0,1],\ p\left(\frac{a^{2}}{2\sigma^{2}}-\beta\right)\leq 0,\ p\left(\frac{a^{2}}{2\sigma^{2}}-\beta\right)+(1-p)\left(\frac{b^{2}}{2}-1\right)\leq 0,

by law of large numbers there should be with high probability particles of type 22 to the right of the position t(pa+(1p)b)t(pa+(1-p)b) at time tt.

If we write (p,a,b)(p^{*},a^{*},b^{*}) the triplet optimizing the problem (2.1), it follows from classical optimization under constraints computations that:

  1. 1.

    If (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, then p=1p^{*}=1 and a=2βσ2a^{*}=\sqrt{2\beta\sigma^{2}}, which is in accordance with Theorem 1.1, as the extremal particle system is dominated by the behaviour of particles of type 11, and particles of type 22 contributing to the extremal process are close relatives descendants of a parent of type 11;

  2. 2.

    If (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, then p=0p^{*}=0 and b=2b^{*}=\sqrt{2}, which is in accordance with Theorem 1.2, as the extremal particle system is dominated by the behaviour of particles of type 22, that are born at time o(t)o(t) from particles of type 11;

  3. 3.

    If (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III}, then

    p=σ2+β22(1σ2)(β1),a=σ22β11σ2 and b=2β11σ2.p^{*}=\frac{\sigma^{2}+\beta-2}{2(1-\sigma^{2})(\beta-1)},\quad a^{*}=\sigma^{2}\sqrt{2\frac{\beta-1}{1-\sigma^{2}}}\ \text{ and }\ b^{*}=\sqrt{2\frac{\beta-1}{1-\sigma^{2}}}.

    We then have v=βσ22(1σ2)(β1)v=\frac{\beta-\sigma^{2}}{\sqrt{2(1-\sigma^{2})(\beta-1)}}, which corresponds to the main result of Theorem 1.3. Additionally, the Lagrange multiplier associated to this optimization problem is θ=2β11σ2=b=aσ2\theta=\sqrt{2\frac{\beta-1}{1-\sigma^{2}}}=b=\frac{a}{\sigma^{2}}.

In particular, the optimization problem associated to the case (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III} can be related to the following interpretation of Theorem 1.3. The extremal process at time tt is obtained as the superposition of the extremal processes of an exponentially large number of BBMs of type 22, starting around time tptp^{*} and position tpatp^{*}a^{*}. The number of these BBMs is directly related to the number of particles of type 11 that displace at speed aa^{*}, which is known to be proportional to W(θ)et(1(a)2/2σ2)W_{\infty}(\theta)e^{t(1-(a^{*})^{2}/2\sigma^{2})}. It explains the apparition of this martingale in Theorem 1.3, whereas the decoration distribution 𝔇(III)\mathfrak{D}^{(III)} is the extremal process of a BBM of type 22 conditionally on moving at the speed b>vb^{*}>v.

For Theorem 1.1, a similar description can be made. We expect the asymptotic behaviour to be driven by the behaviour of particles of type 11, therefore the extremal process of particles of type 22 should be obtained as a decoration of the extremal process of particles of type 11. However, as we were not able to use result of convergence of extremal processes together with a description of the behaviour of particles at times tO(1)t-O(1), we do not obtain an explicit value for c(I)c_{(I)} and an explicit description of the law 𝔇(I)\mathfrak{D}^{(I)}. However, with similar techniques as the ones used in [ABBS13] or [ABK13], such explicit constructions should be available.

Finally, in the case covered by Theorem 1.2, the above optimization problem indicates that the extremal process of the multitype reducible BBM should be obtained as the superposition of a finite number of BBMs of particles of type 22, descending from the first few particles of type 22 to be born. The random variable Z¯\overline{Z} is then constructed as the weighted sum of i.i.d. copies of the derivative martingale of a standard BBM and the decoration is the same as the decoration of the original BBM.

To prove Theorems 1.11.3, we show that the above heuristic holds, i.e. that with high probability the set of particles contributing to the extremal processes are the one we identified in each case. We then use previously known results of branching Brownian motions to compute the Laplace transforms of the extremal point measures we are interested in.

The solution of the optimization problem (2.1) is also solution of v=sup{a:g(a)0}v=\sup\{a\in\mathbb{R}:g(a)\leq 0\}, where gg is the largest convex function such that

|x|2βσ2,g(x)(x22σ2β) and y,g(y)y221,\forall|x|\leq\sqrt{2\beta\sigma^{2}},\ g(x)\leq\left(\frac{x^{2}}{2\sigma^{2}}-\beta\right)\quad\text{ and }\quad\forall y\in\mathbb{R},\ g(y)\leq\frac{y^{2}}{2}-1,

see [Big10] for precisions. The function xx22σ2βx\mapsto\frac{x^{2}}{2\sigma^{2}}-\beta is known as the rate function for particles of type 11, and yy221y\mapsto\frac{y^{2}}{2}-1 is the rate function for particles of type 22.

We then observe that the three cases described above are the following:

  1. 1.

    If (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, then v=2βσ2=sup{x:x2/2σ2β0}v=\sqrt{2\beta\sigma^{2}}=\sup\{x\in\mathbb{R}:x^{2}/2\sigma^{2}-\beta\leq 0\}.

  2. 2.

    If (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, then v=2=sup{y:y2210}v=\sqrt{2}=\sup\{y\in\mathbb{R}:\frac{y^{2}}{2}-1\leq 0\}.

  3. 3.

    If (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III}, then v>max(2βσ2,2)v>\max(\sqrt{2\beta\sigma^{2}},\sqrt{2}).

In other words, the anomalous spreading corresponds to the case when the convex envelope gg crosses the xx-axis to the right of the rate functions of particles of type 11 and 22.

vv
(a) Case I: The speed of the multitype process is the same as the speed of the process of type 1 particles.
vv
(b) Case II: The speed of the multitype process is the same as the speed of the process of type 2 particles.
vv
(c) Case III: The anomalous spreading of the multitype process which is faster than the process consisting only of particles of type 1 or 2.
Figure 2: Convex envelope gg for (β,σ2)(\beta,\sigma^{2}) in any of the three domains of interest. The rate function of particles of type 1 and 2 are drawn in blue and yellow respectively, the function gg of the multitype branching process is drawn in green.

As mentioned above, Ren and Yang [RY14] studied the asymptotic behaviour of irreducible multitype BBM, and computed the speed at which that process invades its environment. In that case (i.e. when for all pair of types ii and jj, individuals of type ii have positive probability of having at least one descendant of type jj after some time), this asymptotic behaviour is similar to the one obtained for a single-type BBM, with branching rate and variance obtained by considering the invariant measure of the Markov process describing the type of a typical individual. The notion of anomalous spreading in this case is thus very different, and the ancestral lineage of typical particles in the extremal process will present regular changes of type. As a result, we do not expect an asymptotic behaviour similar to the one observed in Theorem 1.3 to occur in irreducible multitype BBM.

In a different direction, Smadi and Vatutin [SV16] studied the limit in distribution of a critical reducible Galton-Watson process. It is worth noting that similarly to our results, they obtained three different behaviours for the system, with either the domination of particles of the first type, of the second type, or an interplay between the two.

2.2 Relation to time-inhomogeneous branching processes

The results presented here, in particular in the anomalous spreading case, are reminiscent of the known asymptotic for the extremal process of time-inhomogeneous branching Brownian motions. This model was introduced by Fang and Zeitouni [FZ12a], and is defined as follows. Given t0t\geq 0, the process is a BBM consisting only of particles of type 11 until time t/2t/2, at which time they all become simultaneously particles of type 22. It has been showed [FZ12b, Mal15a] that depending on the value of (β,σ2)(\beta,\sigma^{2}), the position of the maximal displacement at time tt can exhibit different types of asymptotic behaviours. In particular, the logarithmic correction exhibit a strong phase transition in the phase space of (β,σ2)(\beta,\sigma^{2}).

Looking more closely at the convergence of the extremes, Bovier and Hartung [BH14, BH15] obtained the convergence in distribution of the extremal process of the time-inhomogeneous BBM. In particular, for a multitype BBM with parameters (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III} such that particles change from type 11 to type 22 at time ptp^{*}t, they showed that the extremal process converges towards (III)\mathcal{E}^{(III)}_{\infty}, with an extra 12θlogt\frac{1}{2\theta}\log t logarithmic correction for the centring. This is in accordance with our heuristic as we expect that the particles contributing to the extremal process at time tt to have been born from particles of type 11 around time ptp^{*}t.

Generalized versions of time-inhomogeneous BBM have been studied, in which the variance of particles evolves continuously over time [MZ16, Mal15b]. In that case, the maximal displacement grows at constant speed with a negative correction of order t1/3t^{1/3}. It would be interesting to construct a multitype BBM, possibly with an infinite number of types, that would exhibit a similar phenomenon.

2.3 F-KPP type equation associated to the multitype branching Brownian motion

Observe that similarly to the standard BBM, the multitype BBM can be associated to a reaction diffusion equation in the following way. Let f,g:[0,1]f,g:\mathbb{R}\to[0,1] be measurable functions, we define for all xx\in\mathbb{R}:

u(t,x)\displaystyle u(t,x) =𝐄(1)(u𝒩t1f(Xu(t)+x)u𝒩t2g(Xu(t)+x))\displaystyle=\mathbf{E}^{(1)}\left(\prod_{u\in\mathcal{N}^{1}_{t}}f(X_{u}(t)+x)\prod_{u\in\mathcal{N}^{2}_{t}}g(X_{u}(t)+x)\right)
v(t,x)\displaystyle v(t,x) =𝐄(2)(u𝒩t1f(Xu(t)+x)u𝒩t2g(Xu(t)+x))=𝐄(2)(u𝒩t2g(Xu(t)+x))\displaystyle=\mathbf{E}^{(2)}\left(\prod_{u\in\mathcal{N}^{1}_{t}}f(X_{u}(t)+x)\prod_{u\in\mathcal{N}^{2}_{t}}g(X_{u}(t)+x)\right)=\mathbf{E}^{(2)}\left(\prod_{u\in\mathcal{N}^{2}_{t}}g(X_{u}(t)+x)\right)

where 𝐏(1)\mathbf{P}^{(1)} (respectively 𝐏(2)\mathbf{P}^{(2)}) is the law of the multitype BBM starting from one particle of type 11 (resp. 22), and we use the fact that particles of type 22 only produce offspring of type 22, with the usual convention uf(Xu(t)+x)=1\prod_{u\in\emptyset}f(X_{u}(t)+x)=1.

As under 𝐏(2)\mathbf{P}^{(2)}, the process behaves as a standard BBM, the function vv is a solution of the classical F-KPP reaction-diffusion equation

tv=12Δvv(1v)with v(0,x)=g(x).\partial_{t}v=\frac{1}{2}\Delta v-v(1-v)\quad\text{with }v(0,x)=g(x). (2.2)

To obtain the partial differential equation satisfied by uu, we observe that under law 𝐏(1)\mathbf{P}^{(1)} one of the three following events might happen during the first dt\mathrm{d}t units of time:

  • with probability βdt+o(dt)\beta\mathrm{d}t+o(\mathrm{d}t), the original particle of type 11 branches into two offspring of type 11 that start i.i.d. processes with law 𝐏(1)\mathbf{P}^{(1)};

  • with probability αdt+o(dt)\alpha\mathrm{d}t+o(\mathrm{d}t) the particle of type 11 branches into one offspring of type 11 and one of type 22, that start independent processes with law 𝐏(1)\mathbf{P}^{(1)} and 𝐏(2)\mathbf{P}^{(2)} respectively;

  • with probability 1(β+α)dt1-(\beta+\alpha)\mathrm{d}t, the particle of type 22 diffuses as the Brownian motion σB\sigma B with diffusion constant σ2\sigma^{2}

As a result, we have

u(t+dt,x)\displaystyle u(t+\mathrm{d}t,x) =βdtu(t,x)2+αdtu(t,x)v(t,x)+(1(β+α)dt)𝐄(u(t,xσBdt))+o(dt)\displaystyle=\beta\mathrm{d}tu(t,x)^{2}+\alpha\mathrm{d}tu(t,x)v(t,x)+(1-(\beta+\alpha)\mathrm{d}t)\mathbf{E}\left(u(t,x-\sigma B_{\mathrm{d}t})\right)+o(\mathrm{d}t)
=u(t,x)+dt(σ22Δu(t,x)βu(1u)αu(1v)).\displaystyle=u(t,x)+\mathrm{d}t\left(\frac{\sigma^{2}}{2}\Delta u(t,x)-\beta u(1-u)-\alpha u(1-v)\right).

This, together with (2.2) show that (u,v)(u,v) is a solution of the following coupled F-KPP equation

{tu=σ22Δβu(1u)αu(1v)tv=12Δvv(1v)u(0,x)=f(x),v(0,x)=g(x).\begin{cases}&\partial_{t}u=\frac{\sigma^{2}}{2}\Delta-\beta u(1-u)-\alpha u(1-v)\\ &\partial_{t}v=\frac{1}{2}\Delta v-v(1-v)\\ &u(0,x)=f(x),\quad v(0,x)=g(x).\end{cases} (2.3)

This non-linear coupling of F-KPP equation was introduced by Holzer [Hol14]. In that article, the author conjectured this partial differential equation to exhibit an anomalous spreading phenomenon, and conjectured a phase diagram for the model [Hol14, Figure 1]. Our main results confirm this conjecture, and the diagram we obtain in Figure 1 exactly matches (up to an adaptation of the notation σ22d\sigma^{2}\rightsquigarrow 2d, βα\beta\rightsquigarrow\alpha and αβ\alpha\rightsquigarrow\beta) the one obtained by Holzer. Additionally, Theorems 1.11.3 give the position of the front of vtv_{t} in (2.3).

When starting with well chosen initial conditions ff,gg (for example such that there exists A>0A>0 satisfying f(x)=g(x)=1f(x)=g(x)=1 for x<Ax<-A and f(x)=g(x)=0f(x)=g(x)=0 for x>Ax>A), we obtain the existence of a function vtv_{t} such that for all xx\in\mathbb{R},

limt(u(t,xvt),v(t,xvt))=(w1(x),w2(x)),\lim_{t\to\infty}(u(t,x-v_{t}),v(t,x-v_{t}))=(w_{1}(x),w_{2}(x)),

where (w1,w2)(w_{1},w_{2}) is a travelling wave solution of the coupled PDE and:

  1. 1.

    if (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, then vt=2βσ2t322β/σ2logtv_{t}=\sqrt{2\beta\sigma^{2}}t-\frac{3}{2\sqrt{2\beta/\sigma^{2}}}\log t;

  2. 2.

    if (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, then vt=2t322logtv_{t}=\sqrt{2}t-\frac{3}{2\sqrt{2}}\log t;

  3. 3.

    if (β,σ2)𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{III}, then vt=vtv_{t}=vt, with vv defined in Theorem 1.3.

Holzer further studied a linearised version of (2.3) in [Hol16], and showed the presence of an anomalous spreading property in that context. However, the phase diagram in that case is of a different nature as the one we obtain in Figure 1. We believe that the phase diagram of this linearised PDE equation should be related to first moment estimates on the number of particles above a given level in the multitype BBM.

The equation (2.3) should also be compared to the partial differential equation studied in [BC14]. In that article, they considered a population with a family of traits indexed by a parameter θ(θmin,θmax)\theta\in(\theta_{\min},\theta_{\max}), that modifies the motility of particles. This was proposed as a model for the invasion of cane toads is Australia, as that population consists of faster individuals, that sacrifice part of their reproduction power as a trade off, and slower individuals that reproduce more easily. The multitype BBM we consider here could then be thought of as some toy-model for this partial differential equation.

2.4 Future developments

We recall that Theorems 1.11.3 cover the asymptotic behaviour of the two-type reducible BBM assuming that (β,σ2)𝒞I𝒞II𝒞III(\beta,\sigma^{2})\in\mathcal{C}_{I}\cup\mathcal{C}_{II}\cup\mathcal{C}_{III}. However, it does not give the asymptotic behaviour of this process when (β,σ2)(\beta,\sigma^{2}) belongs to the boundary of this set. Understanding the behaviour of the process at these points could help understanding the phase transitions occurring between the different areas of the state space. This would allow results similar to the ones developed in [BH20] for time-inhomogeneous BBM to be considered in reducible multitype BBM.

We conjecture the following behaviours for the branching Brownian motion at the boundary between areas 𝒞I\mathcal{C}_{I} and 𝒞III\mathcal{C}_{III}.

Conjecture 2.1.

Assume that β>1\beta>1 and σ2=β2β1\sigma^{2}=\frac{\beta}{2\beta-1}, then there exist c>0c>0 and 𝔇~\widetilde{\mathfrak{D}} such that

u𝒩t2δXu(t)2βσ2t+12β/σ2logt\sum_{u\in\mathcal{N}_{t}^{2}}\delta_{X_{u}(t)-\sqrt{2\beta\sigma^{2}}t+\frac{1}{\sqrt{2\beta/\sigma^{2}}}\log t}

converges to a DPPP(cZ(1)e2β/σ2xdx,𝔇~cZ^{(1)}_{\infty}e^{-\sqrt{2\beta/\sigma^{2}}x}\mathrm{d}x,\widetilde{\mathfrak{D}}).

Indeed, in this situation, particles uu of type 22 contributing to the extremal process are expected to satisfy tT(u)=O(t1/2)t-T(u)=O(t^{1/2}). Therefore, the extremal process keeps an intensity driven by the derivative martingale of particles of type 11, and the decoration point measure is given by the extremal process of a BBM of particles of type 22 conditioned to travel at speed 2βσ2>2\sqrt{2\beta\sigma^{2}}>\sqrt{2}.

Similarly, at the boundary between areas 𝒞II\mathcal{C}_{II} and 𝒞III\mathcal{C}_{III}, the following behaviour is expected.

Conjecture 2.2.

Assume that β>1\beta>1 and σ2=2β\sigma^{2}=2-\beta, then there exist c>0c>0 and a random variable Z~\widetilde{Z} such that

u𝒩t2δXu(t)2t+12logt\sum_{u\in\mathcal{N}_{t}^{2}}\delta_{X_{u}(t)-\sqrt{2}t+\frac{1}{\sqrt{2}}\log t}

converges to a DPPP(cZ~e2xdx,𝔇c\widetilde{Z}e^{-\sqrt{2}x}\mathrm{d}x,{\mathfrak{D}}).

There, we used the fact that particles uu of type 22 contributing to the extremal process are expected to satisfy T(u)=O(t1/2)T(u)=O(t^{1/2}).

In the case when β<1\beta<1 and σ2β=1\sigma^{2}\beta=1, which corresponds to the boundary between cases 𝒞I\mathcal{C}_{I} and 𝒞II\mathcal{C}_{II}, the picture is less clear as at all time ss between 0 and tt, particles should have the same probability to reach the maximal position, at least to the first order, as the BBM of particles of type 11 and of particles of type 22 have same speed.

Further generalisations of the model we consider in this article could be considered. A more general reducible multitype branching Brownian motions with a finite number of states would be expected to exhibit a similar behaviour. One could also allow particles to have different drift coefficients in addition to the different variance terms and branching rates. In that situation, one expects an optimization problem similar to the one studied in [Mal15a] to appear, with a similar resolution of proving that the trajectory followed by particles reaching the maximal position is the same as the one inferred from the solution of the optimization problem.

Proving Theorems 1.11.3 for two-type reducible branching Brownian motions in which particles of type 11 and type 22 split into a random number of children at each branching event, say L1L_{1} for particles of type 11 and L2L_{2} for particles of type 22 would be a other natural generalisation of our results. A natural condition to put on the reproduction laws to obtain the asymptotic behaviour observed in Theorem 1.3 is

𝐄(L1logL1)+𝐄(L2logL2)<.\mathbf{E}(L_{1}\log L_{1})+\mathbf{E}(L_{2}\log L_{2})<\infty.

It is worth noting that anomalous spreading might occur even if 𝐄(L2)<1\mathbf{E}(L_{2})<1, i.e. even if the genealogical tree of a particle of type 22 is subcritical and grows extinct almost surely.

While we only take interest here in the asymptotic behaviour of the extremal particles in this article, we believe that many other features of multitype branching Brownian motions might be of interest, such as the growth rate of the number of particles of type 22 to the right of atat for a<va<v, the large deviations of the maximal displacement MtM_{t} at time tt, or the convergence of associated (sub)-martingales.

3 Preliminary results on the branching Brownian motion

We list in this section results on the standard BBM, that we use to study the two-type reducible BBM. For the rest of the section, (Xu(t),u𝒩t)t0(X_{u}(t),u\in\mathcal{N}_{t})_{t\geq 0} will denote a standard BBM, with branching rate 11 and diffusion constant 11, i.e. that has the same behaviour as particles of type 22. To translate the results of this section to the behaviour of particles of type 11 as well, it is worth noting that for all β,σ>0\beta,\sigma>0:

(σβXu(βt),u𝒩βt)t0\left(\frac{\sigma}{\sqrt{\beta}}X_{u}(\beta t),u\in\mathcal{N}_{\beta t}\right)_{t\geq 0} (3.1)

is a branching Brownian with branching rate β\beta and diffusion constant σ2\sigma^{2}.

The rest of the section is organised as follows. We introduce in Section 3.1 the additive martingales of the BBM, and in particular the derivative martingale that plays a special role in the asymptotic behaviour of the maximal displacement of the BBM. We then provide in Section 3.3 a series of uniform asymptotic estimates on the maximal displacement of the BBM. Finally, in Section 3.4, we introduce the decoration measures and extremal processes appearing when studying particles near the rightmost one in the BBM.

3.1 Additive martingales of the branching Brownian motion

We begin by introducing the additive martingales of the BBM. For all θ\theta\in\mathbb{R}, the process

Wt(θ):=u𝒩teθXu(t)t(θ22+1),t0W_{t}(\theta):=\sum_{u\in\mathcal{N}_{t}}e^{\theta X_{u}(t)-t\left(\tfrac{\theta^{2}}{2}+1\right)},\quad t\geq 0 (3.2)

is a non-negative martingale. It is now a well-known fact that the martingale (Wt(θ),t0)(W_{t}(\theta),t\geq 0) is uniformly integrable if and only if |θ|<2|\theta|<\sqrt{2}, and in that case it converges towards an a.s. positive limiting random variable

W(θ):=limtWt(θ).W_{\infty}(\theta):=\lim_{t\to\infty}W_{t}(\theta). (3.3)

Otherwise, we have limtWt(θ)=0\lim_{t\to\infty}W_{t}(\theta)=0 a.s. This result was first shown by [Nev88]. It can also be obtained by a specific change of measure technique, called the spinal decomposition. This method was pioneered by Lyons, Pemantle and Peres [LPP95] for the study of the martingale of a Galton-Watson process, and extended by Lyons [Lyo97] to spatial branching processes setting.

For all |θ|<2|\theta|<\sqrt{2} the martingale limit W(θ)W_{\infty}(\theta) is closely related to the number of particles moving at speed θ\theta in the BBM. For example, by [Big92, Corollary 4], for all h>0h>0 we have

limtt1/2et(θ221)u𝒩t𝟙{|Xu(t)θt|h}=2sinh(θh)θW(θ)a.s.\lim_{t\to\infty}t^{1/2}e^{t\left(\frac{\theta^{2}}{2}-1\right)}\sum_{u\in\mathcal{N}_{t}}\mathbbm{1}_{\left\{|X_{u}(t)-\theta t|\leq h\right\}}=\frac{2\sinh(\theta h)}{\theta}W_{\infty}(\theta)\quad\text{a.s.} (3.4)

This can be thought of as a local limit theorem result for the position of a particle sampled at random at time tt, where a particle at position xx is sampled with probability proportional to eθxe^{\theta x}. A Donsker-type theorem was obtained in [Pai18, Section C] for this quantity, see also [GKS18]. In particular, for any continuous bounded function ff, one has

limtu𝒩tf(Xu(t)θtt1/2)eθXu(t)t(θ22+1)=W(θ)2πez22f(z)dz a.s.\lim_{t\to\infty}\sum_{u\in\mathcal{N}_{t}}f\left(\tfrac{X_{u}(t)-\theta t}{t^{1/2}}\right)e^{\theta X_{u}(t)-t\left(\frac{\theta^{2}}{2}+1\right)}=\frac{W_{\infty}(\theta)}{\sqrt{2\pi}}\int_{\mathbb{R}}e^{-\frac{z^{2}}{2}}f(z)\mathrm{d}z\text{ a.s.} (3.5)

This justifies the fact that the variable W(θ)W_{\infty}(\theta) appears in the limiting distribution of the extremal process in the anomalous spreading case, by the heuristics described in Section 2.1.

To prove Theorem 1.3, we use the following slight generalization of the above convergence.

Lemma 3.1.

Let a<ba<b and λ>0\lambda>0. For all continuous bounded function f:[a,b]×f:[a,b]\times\mathbb{R}\to\mathbb{R}, we have

limt1t1/2λt+at1/2λt+bt1/2u𝒩sf(sλtt1/2,Xu(s)θst1/2)eθXu(s)s(θ22+1)ds=W(θ)2πλ[a,b]×ez22λf(r,z)drdz a.s.\lim_{t\to\infty}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}f\left(\tfrac{s-\lambda t}{t^{1/2}},\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s=\frac{W_{\infty}(\theta)}{\sqrt{2\pi\lambda}}\int_{[a,b]\times\mathbb{R}}e^{-\frac{z^{2}}{2\lambda}}f(r,z)\mathrm{d}r\mathrm{d}z\text{ a.s.} (3.6)
Proof.

As a first step, we show that (3.6) holds for f:(r,x)𝟙{r[a,b]}g(x)f:(r,x)\mapsto\mathbbm{1}_{\left\{r\in[a,b]\right\}}g(x), with gg a continuous compactly supported function. Using that

limsu𝒩sg(λ1/2Xu(s)θss1/2)eθXu(s)s(θ22+1)=W(θ)2πez22g(λ1/2z)dza.s,\lim_{s\to\infty}\sum_{u\in\mathcal{N}_{s}}g\left(\lambda^{1/2}\tfrac{X_{u}(s)-\theta s}{s^{1/2}}\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}=\frac{W_{\infty}(\theta)}{\sqrt{2\pi}}\int_{\mathbb{R}}e^{-\frac{z^{2}}{2}}g(\lambda^{1/2}z)\mathrm{d}z\quad\text{a.s,}

by (3.5) we immediately obtain that

limt1t1/2λt+at1/2λt+bt1/2u𝒩sg(λ1/2Xu(s)θss1/2)eθXu(s)s(θ22+1)ds=(ba)W(θ)2πλez22λg(z)dz a.s.\lim_{t\to\infty}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}g\left(\lambda^{1/2}\tfrac{X_{u}(s)-\theta s}{s^{1/2}}\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s=(b-a)\frac{W_{\infty}(\theta)}{\sqrt{2\pi\lambda}}\int_{\mathbb{R}}e^{-\frac{z^{2}}{2\lambda}}g(z)\mathrm{d}z\text{ a.s.}

We then observe that for all s[λt+at1/2,λt+bt1/2]s\in[\lambda t+at^{1/2},\lambda t+bt^{1/2}], we have

s1/2=(λt+sλt)1/2=(λt)1/2(1+sλtλt)1/2.s^{1/2}=\left(\lambda t+s-\lambda t\right)^{1/2}=(\lambda t)^{1/2}\left(1+\frac{s-\lambda t}{\lambda t}\right)^{1/2}.

As sλtλt[at1/2/λ,bt1/2/λ]\frac{s-\lambda t}{\lambda t}\in[at^{-1/2}/\lambda,bt^{-1/2}/\lambda], there exists a constant K>0K>0 such that for all t1t\geq 1, we have

sups[λt+at1/2,λt+bt1/2]|(1+sλtλt)1/21|Kt1/2,\sup_{s\in[\lambda t+at^{1/2},\lambda t+bt^{1/2}]}\left|\left(1+\frac{s-\lambda t}{\lambda t}\right)^{1/2}-1\right|\leq Kt^{-1/2},

so |s1/2(λt)1/2|Kλ1/2\left|s^{1/2}-(\lambda t)^{1/2}\right|\leq K\lambda^{1/2} uniformly in s[λt+at1/2,λt+bt1/2]s\in[\lambda t+at^{1/2},\lambda t+bt^{1/2}], for all tt large enough.

Then, using the uniform continuity and compactness of gg, for all ε>0\varepsilon>0 we have

sups[λt+at1/2,λt+bt1/2b],x|g(λ1/2x/s1/2)g(x/t1/2)|ε\sup_{s\in[\lambda t+at^{1/2},\lambda t+bt^{1/2}b],x\in\mathbb{R}}\left|g(\lambda^{1/2}x/s^{1/2})-g(x/t^{1/2})\right|\leq\varepsilon

for all tt large enough. Therefore, for all ε>0\varepsilon>0,

lim supt1t1/2λt+at1/2λt+bt1/2u𝒩s|g(Xu(s)θss1/2)g(Xu(s)θs(λt)1/2)|eθXu(s)s(θ22+1)dslim supt1t1/2λt+at1/2λt+bt1/2u𝒩sεeθXu(s)s(θ22+1)ds=ε(ba)W(θ)a.s.\limsup_{t\to\infty}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}\left|g\left(\tfrac{X_{u}(s)-\theta s}{s^{1/2}}\right)-g\left(\tfrac{X_{u}(s)-\theta s}{(\lambda t)^{1/2}}\right)\right|e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s\\ \leq\limsup_{t\to\infty}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}\varepsilon e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s=\varepsilon(b-a)W_{\infty}(\theta)\quad\text{a.s.}

Letting ε0\varepsilon\to 0, we finally obtain

limt1t1/2λt+at1/2λt+bt1/2u𝒩sf(sλtt1/2,Xu(s)θst1/2)eθXu(s)s(θ22+1)ds=W(θ)2πλ[a,b]×ez22λf(r,z)drdza.s.\lim_{t\to\infty}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}f\left(\tfrac{s-\lambda t}{t^{1/2}},\tfrac{X_{u}(s)-\theta^{s}}{t^{1/2}}\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s=\frac{W_{\infty}(\theta)}{\sqrt{2\pi\lambda}}\int_{[a,b]\times\mathbb{R}}\!\!\!e^{-\frac{z^{2}}{2\lambda}}f(r,z)\mathrm{d}r\mathrm{d}z\quad\text{a.s.} (3.7)

We now assume that ff is a continuous compactly supported function on [a,b]×[a,b]\times\mathbb{R}. For all ini\leq n, we set

fi(r,x)=𝟙{r[a+i(ba)/n,a+(i+1)(ba)/n]}f(a+i(ba)/n,x).f_{i}(r,x)=\mathbbm{1}_{\left\{r\in[a+i(b-a)/n,a+(i+1)(b-a)/n]\right\}}f(a+i(b-a)/n,x).

Using the uniform integrability of ff, for all nn large enough, we have fj=1nfjε{\left\|f-\sum_{j=1}^{n}f_{j}\right\|}_{\infty}\leq\varepsilon. As a result, we have

lim supt1t1/2λt+at1/2λt+bt1/2u𝒩s|f(sλtt1/2,Xu(s)θst1/2)i=1nfi(sλtt1/2,Xu(s)θst1/2)|eθXu(s)s(θ22+1)dsεW(θ) a.s.\limsup_{t\to\infty}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}\left|f\left(\tfrac{s-\lambda t}{t^{1/2}},\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)-\sum_{i=1}^{n}f_{i}\left(\tfrac{s-\lambda t}{t^{1/2}},\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)\right|e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s\leq\varepsilon W_{\infty}(\theta)\text{ a.s.}

Therefore, using (3.7), we obtain

lim supt|1t1/2λt+at1/2λt+bt1/2u𝒩sf(sλtt1/2,Xu(s)θst1/2)eθXu(s)s(θ22+1)dsW(θ)2πλ[a,b]×ez22λf(r,z)drdz|2εW(θ)a.s.\limsup_{t\to\infty}\Bigg{|}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}f\left(\tfrac{s-\lambda t}{t^{1/2}},\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s\\ -\frac{W_{\infty}(\theta)}{\sqrt{2\pi\lambda}}\int_{[a,b]\times\mathbb{R}}\!\!\!e^{-\frac{z^{2}}{2\lambda}}f(r,z)\mathrm{d}r\mathrm{d}z\Bigg{|}\leq 2\varepsilon W_{\infty}(\theta)\quad\text{a.s.}

Letting ε0\varepsilon\to 0 therefore proves that (3.6) holds for compactly supported continuous functions.

Finally, to complete the proof, we consider a continuous bounded function ff on [a,b]×[a,b]\times\mathbb{R}. Let R>0R>0, given χR\chi_{R} a continuous function on \mathbb{R} such that 𝟙{|x|<R}χR(x)𝟙{|x|R+1}\mathbbm{1}_{\left\{|x|<R\right\}}\leq\chi_{R}(x)\leq\mathbbm{1}_{\left\{|x|\leq R+1\right\}}, the previous computation shows that

limt1t1/2λt+at1/2λt+bt1/2u𝒩sχR(Xu(s)θst1/2)f(sλtt1/2,Xu(s)θst1/2)eθXu(s)s(θ22+1)ds=W(θ)2πλ[a,b]×ez22λf(r,z)χR(z)drdz a.s.\lim_{t\to\infty}\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}\chi_{R}\left(\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)f\left(\tfrac{s-\lambda t}{t^{1/2}},\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s\\ =\frac{W_{\infty}(\theta)}{\sqrt{2\pi\lambda}}\int_{[a,b]\times\mathbb{R}}e^{-\frac{z^{2}}{2\lambda}}f(r,z)\chi_{R}(z)\mathrm{d}r\mathrm{d}z\text{ a.s.}

Additionally, setting K=fK={\left\|f\right\|}_{\infty}, for all tt large enough we have

|1t1/2λt+at1/2λt+bt1/2u𝒩s(1χR(Xu(s)θst1/2))f(sλtt1/2,Xu(s)θst1/2)eθXu(s)s(θ22+1)ds|Kt1/2λt+at1/2λt+bt1/2u𝒩s(1χR(λ1/2Xu(s)θs2s1/2))eθXu(s)s(θ22+1)ds,\left|\frac{1}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}\left(1-\chi_{R}\left(\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)\right)f\left(\tfrac{s-\lambda t}{t^{1/2}},\tfrac{X_{u}(s)-\theta s}{t^{1/2}}\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s\right|\\ \leq\frac{K}{t^{1/2}}\int_{\lambda t+at^{1/2}}^{\lambda t+bt^{1/2}}\sum_{u\in\mathcal{N}_{s}}\left(1-\chi_{R}\left(\lambda^{1/2}\tfrac{X_{u}(s)-\theta s}{2s^{1/2}}\right)\right)e^{\theta X_{u}(s)-s\left(\frac{\theta^{2}}{2}+1\right)}\mathrm{d}s,

which converges to (ba)W(θ)2π(1χR(λ1/2z/2))ez22dz\frac{(b-a)W_{\infty}(\theta)}{\sqrt{2\pi}}\int_{\mathbb{R}}(1-\chi_{R}(\lambda^{1/2}z/2))e^{-\frac{z^{2}}{2}}\mathrm{d}z as tt\to\infty. Thus, letting tt\to\infty then RR\to\infty completes the proof of this lemma. ∎

3.2 The derivative martingale

The number of particles that travel at the critical speed 2\sqrt{2} cannot be counted using the additive martingale (as it converges to 0 almost surely). In this situation, the appropriate process allowing this estimation is the derivative martingale (Zt,t0)(Z_{t},t\geq 0). Its name comes from the fact that ZtZ_{t} can be represented as θWt(θ)|θ=2\left.-\frac{\partial}{\partial\theta}W_{t}(\theta)\right|_{\theta=\sqrt{2}}, more precisely

Zt:=u𝒩t(2tXu(t))e2Xu(t)2t.Z_{t}:=\sum_{u\in\mathcal{N}_{t}}(\sqrt{2}t-X_{u}(t))e^{\sqrt{2}X_{u}(t)-2t}. (3.8)

Despite being a non-integrable signed martingale, it was proved by Lalley and Sellke [LS87] that it converges to an a.s. positive random variable

Z:=limtZt a.s.Z_{\infty}:=\lim_{t\to\infty}Z_{t}\quad\text{ a.s.} (3.9)

In the same way that the limit of the additive martingale gives the growth rate of the number of particles moving at speed θ\theta, the derivative martingale gives the growth rate of particles that go at speed 2\sqrt{2}. As a result, it appears in the asymptotic behaviour of the maximal displacement, and results similar to (3.4) and (3.5) can be found in [Mad16, Pai18] in the context of branching random walks.

We mention that the limit ZZ_{\infty} of the derivative martingale is non-integrable, and that its precise tail has been well-studied. In particular, Bereskycki, Berestycki and Schweinsberg [BBS13] proved that

𝐏(Zx)2x as x.\mathbf{P}(Z_{\infty}\geq x)\sim\frac{\sqrt{2}}{x}\text{ as }x\to\infty. (3.10)

Similar results were obtained for branching random walks by Buraczewski [Bur09] and Madaule [Mad16]. They also obtained a more precise estimate on its asymptotic, that can be expressed in the two following equivalent ways

𝐄(Z𝟙{Zx})=2logx+O(1) as x,\displaystyle\mathbf{E}(Z_{\infty}\mathbbm{1}_{\left\{Z_{\infty}\leq x\right\}})=\sqrt{2}\log x+O(1)\quad\text{ as }x\to\infty, (3.11)
1𝐄(eλZ)=2λlogλ+O(λ) as λ0.\displaystyle 1-\mathbf{E}(e^{-\lambda Z_{\infty}})=\sqrt{2}\lambda\log\lambda+O(\lambda)\quad\text{ as }\lambda\to 0. (3.12)

Maillard and Pain [MP19] improved on these statements and gave necessary and sufficient conditions for the asymptotic developments of these quantities up to a o(1)o(1). We mention that the equivalence between (3.11) and (3.12) can be found in [BIM20, Lemma 8.1], which obtain similar necessary and sufficient conditions for the asymptotic development of the Laplace transform of the derivative martingale of the branching random walk under optimal integrability conditions.

3.3 Maximal displacement of the branching Brownian motion

A large body of work has been dedicated to the study of the maximal displacement of the BBM, defined by Mt=maxu𝒩tXu(t)M_{t}=\max_{u\in\mathcal{N}_{t}}X_{u}(t). We recall here some estimates related to its study. We begin by observing that the BBM travels in a triangular-shaped array, and that for all y0y\geq 0

𝐏(t0,u𝒩t:Xu(t)2t+y)e2y,\mathbf{P}\left(\exists t\geq 0,u\in\mathcal{N}_{t}:X_{u}(t)\geq\sqrt{2}t+y\right)\leq e^{-\sqrt{2}y}, (3.13)

which shows that with high probability, all particles at time tt are smaller than 2t+y\sqrt{2}t+y in absolute value.

Recall that Lalley and Sellke [LS87] proved that setting mt=2t322logtm_{t}=\sqrt{2}t-\frac{3}{2\sqrt{2}}\log t, the maximal displacement of the BBM centered by mtm_{t} converges in distribution to a shifted Gumbel distribution. More precisely, there exists c>0c_{\star}>0 such that

limt𝐏(Mtmt+z)=𝐄(exp(cZe2z)).\lim_{t\to\infty}\mathbf{P}(M_{t}\leq m_{t}+z)=\mathbf{E}\left(\exp\left(-c_{\star}Z_{\infty}e^{-\sqrt{2}z}\right)\right). (3.14)

An uniform upper bound is also known for the right tail of the maximal displacement. There exists C>0C>0 such that for all t0t\geq 0 and xx\in\mathbb{R}, we have

𝐏(Mtmt+x)C(1+x+)e2x,\mathbf{P}(M_{t}\geq m_{t}+x)\leq C(1+x_{+})e^{-\sqrt{2}x}, (3.15)

where x+=max(x,0)x_{+}=\max(x,0). This estimate can be obtained by first moment methods, we refer e.g. to [Hu16] for a similar estimate in the branching random walk, which immediately implies a similar bound for the BBM.

In the context of the anomalous spreading, seen from the heuristics in Section 2.1, it will also be necessary to use tight estimates on the large deviations of the BBM. These large deviations were first studied by Chauvin and Rouault [CR88]. Precise large deviations for the maximal displacement were recently obtained in [DMS16, GH18, BM19, BBCM20], proving that for all ϱ>2\varrho>\sqrt{2}, there exists C(ϱ)(0,1)C(\varrho)\in(0,1) such that

𝐏(Mtϱt+y)tC(ϱ)2πtϱe(ϱ2/21)teϱyy22t,\mathbf{P}(M_{t}\geq\varrho t+y)\sim_{t\to\infty}\frac{C(\varrho)}{\sqrt{2\pi t}\varrho}e^{-(\varrho^{2}/2-1)t}e^{-\varrho y-\tfrac{y^{2}}{2t}}, (3.16)

uniformly in |y|rt|y|\leq r_{t}, for all function rt=o(t)r_{t}=o(t).

Additionally, from a simple first moment estimate, one can obtain an uniform upper bound for this large deviations estimate on the maximal displacement.

Lemma 3.2.

For all ϱ>2\varrho>\sqrt{2} and A>0A>0, there exists C>0C>0 such that for all tt large enough and all yAt1/2y\geq-At^{1/2}, we have

𝐏(Mtϱt+y)Ce(ϱ2/21)tt1/2eϱyy22t.\mathbf{P}(M_{t}\geq\varrho t+y)\leq\frac{Ce^{-(\varrho^{2}/2-1)t}}{t^{1/2}}e^{-\varrho y-\tfrac{y^{2}}{2t}}.

This result is based on Markov inequality and classical Gaussian estimates, that appear later in our paper in more complicate settings. We thus give a short proof of this statement.

Proof.

Observe that for tt large enough, we have ϱt+yδt\varrho t+y\geq\delta t for some positive constant δ\delta. Then, by Markov inequality, we have

𝐏(Mtϱt+y)=𝐏(u𝒩t:Xu(t)ϱt+y)𝐄(u𝒩t𝟙{Xu(t)ϱt+y}).\mathbf{P}(M_{t}\geq\varrho t+y)=\mathbf{P}\left(\exists u\in\mathcal{N}_{t}:X_{u}(t)\geq\varrho t+y\right)\\ \leq\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}}\mathbbm{1}_{\left\{X_{u}(t)\geq\varrho t+y\right\}}\right).

Using that there are on average ete^{t} particles alive at time tt and that the displacements of particles are Brownian motions, that are independent of the total number of particles in the process, we have

𝐄(u𝒩t𝟙{Xu(t)ϱt+y})=et𝐏(Btϱt+y).\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}}\mathbbm{1}_{\left\{X_{u}(t)\geq\varrho t+y\right\}}\right)=e^{t}\mathbf{P}(B_{t}\geq\varrho t+y).

This fact is often called the many-to-one lemma in the literature (see e.g. [Shi15, Theorem 1]). We develop in Section 4 a multitype versions of this result.

We now use the following well-known asymptotic estimate on the tail of the Gaussian random variable that

𝐏(B1x)12πxex22 for all x0.\mathbf{P}(B_{1}\geq x)\leq\frac{1}{\sqrt{2\pi}x}e^{-\tfrac{x^{2}}{2}}\quad\text{ for all $x\geq 0$}. (3.17)

This yields

𝐄(u𝒩t𝟙{Xu(t)ϱt+y})\displaystyle\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}}\mathbbm{1}_{\left\{X_{u}(t)\geq\varrho t+y\right\}}\right) =et𝐏(B1ϱt1/2+yt1/2)Ct1/2e(ϱt+y)22t\displaystyle=e^{t}\mathbf{P}\left(B_{1}\geq\varrho t^{1/2}+yt^{-1/2}\right)\leq Ct^{-1/2}e^{-\tfrac{(\varrho t+y)^{2}}{2t}}
Ct1/2et(1ϱ2/2)eϱyy22t\displaystyle\leq Ct^{-1/2}e^{t(1-\varrho^{2}/2)}e^{-\varrho y-\tfrac{y^{2}}{2t}}

completing the proof. ∎

3.4 Decorations of the branching Brownian motion

We now turn to results related to the extremal process of the BBM. Before stating these, we introduce a general tool that allows the obtention of the joint convergence in distribution of the maximal displacement and the extremal process of a particle system. Denote by 𝒯\mathcal{T} the set of continuous non-negative bounded functions, with support bounded on the left. The following result can be found in [BBCM20, Lemma 4.4].

Proposition 3.3.

Let 𝒫n,𝒫\mathcal{P}_{n},\mathcal{P} be point measures on the real line. We denote by max𝒫n\max\mathcal{P}_{n} (respectively max𝒫\max\mathcal{P}), the position of the rightmost atom in this point measure. The following statements are equivalent

  1. 1.

    limn𝒫n=𝒫\lim_{n\to\infty}\mathcal{P}_{n}=\mathcal{P} and limnmax𝒫n=max𝒫\lim_{n\to\infty}\max\mathcal{P}_{n}=\max\mathcal{P} in law.

  2. 2.

    limn(𝒫n,max𝒫n)=(𝒫,max𝒫)\lim_{n\to\infty}(\mathcal{P}_{n},\max\mathcal{P}_{n})=(\mathcal{P},\max\mathcal{P}) in law.

  3. 3.

    for all φ𝒯\varphi\in\mathcal{T}, limn𝐄(e𝒫n,φ)=𝐄(e𝒫,φ)\lim_{n\to\infty}\mathbf{E}\left(e^{-{\left\langle\mathcal{P}_{n},\varphi\right\rangle}}\right)=\mathbf{E}\left(e^{-{\left\langle\mathcal{P},\varphi\right\rangle}}\right).

In other words, considering continuous bounded functions with support bounded on the left instead of continuous compactly supported functions allow us to capture the joint convergence in law of the maximal displacement and the extremal process. We refer to the set 𝒯\mathcal{T} as the set of test functions, against which we test the convergence of our point measures of interest.

The convergence in distribution of the extremal process of a BBM has been obtained by Aïdékon, Berestycki, Brunet and Shi [ABBS13], and by Arguin, Bovier and Kistler [ABK13]. They proved that setting

t=u𝒩tδXu(t)mt,\mathcal{E}_{t}=\sum_{u\in\mathcal{N}_{t}}\delta_{X_{u}(t)-m_{t}},

this extremal process converges in distribution towards a decorated Poisson point process with intensity cZ2e2zdzc_{\star}Z_{\infty}\sqrt{2}e^{-\sqrt{2}z}\mathrm{d}z. The law of the decoration is described in [ABK13] as the limiting distribution of the maximal displacement seen from the rightmost particle, conditioned on being larger than 2t\sqrt{2}t at time tt. More precisely, they proved that there exists a point measure 𝒟\mathcal{D} such that

limt𝐄(exp(u𝒩tφ(Xu(t)Mt))|Mt2t)=𝐄(exp(𝒟,φ))\lim_{t\to\infty}\mathbf{E}\left(\exp\left(-\sum_{u\in\mathcal{N}_{t}}\varphi(X_{u}(t)-M_{t})\right)\middle|M_{t}\geq\sqrt{2}t\right)=\mathbf{E}\left(\exp\left(-{\left\langle\mathcal{D},\varphi\right\rangle}\right)\right) (3.18)

for all function φ𝒯\varphi\in\mathcal{T}. Note that 𝒟\mathcal{D} is supported on (,0](-\infty,0] and has an atom at 0.

The limiting extremal process \mathcal{E}_{\infty} can be constructed as follows. Let (ξj)j(\xi_{j})_{j\in\mathbb{N}} be the atoms of a Poisson point process with intensity c2e2zdzc_{\star}\sqrt{2}e^{-\sqrt{2}z}\mathrm{d}z, and (𝒟j,j)(\mathcal{D}_{j},j\in\mathbb{N}) i.i.d. point measures, then set

=jd𝒟jδξj+d+12logZ,\mathcal{E}_{\infty}=\sum_{j\in\mathbb{N}}\sum_{d\in\mathcal{D}_{j}}\delta_{\xi_{j}+d+\frac{1}{\sqrt{2}}\log Z_{\infty}},

where d𝒟j\sum_{d\in\mathcal{D}_{j}} represents a sum on the set of atoms of the point measure 𝒟j\mathcal{D}_{j}.

In view of Proposition 3.3 and (3.14), we can rewrite as follows the convergence in law of the extremal process of the BBM, with simple Poisson computations.

Lemma 3.4.

For all function φ𝒯\varphi\in\mathcal{T}, we have

limt𝐄(et,φ)=𝐄(exp(cZ(1eΨ[φ](z))2e2zdz)),\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\mathcal{E}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(\exp\left(-c_{\star}Z_{\infty}\int(1-e^{-\Psi[\varphi](z)})\sqrt{2}e^{-\sqrt{2}z}\mathrm{d}z\right)\right),

where we have set Ψ[φ]:zlog𝐄(e𝒟,φ(+z))\Psi[\varphi]:z\mapsto-\log\mathbf{E}\left(e^{-{\left\langle\mathcal{D},\varphi(\cdot+z)\right\rangle}}\right).

In the context of large deviations of BBM, a one-parameter family of point measures, similar to the one defined in (3.18) can be introduced. These point measures have first been studied by Bovier and Hartung [BH14] when considering the extremal process of the time-inhomogeneous BBM. More precisely, they proved that for all ϱ>2\varrho>\sqrt{2}, there exists a point measure 𝒟ϱ\mathcal{D}^{\varrho} such that

limt𝐄(exp(u𝒩tφ(Xu(t)Mt))|Mtϱt)=𝐄(exp(𝒟ϱ,φ)).\lim_{t\to\infty}\mathbf{E}\left(\exp\left(-\sum_{u\in\mathcal{N}_{t}}\varphi(X_{u}(t)-M_{t})\right)\middle|M_{t}\geq\varrho t\right)=\mathbf{E}\left(\exp\left(-{\left\langle\mathcal{D}^{\varrho},\varphi\right\rangle}\right)\right). (3.19)

In [BBCM20], an alternative construction of this one parameter family of point measures was introduced, which allows its representation as a point measure conditioned on an event of positive probability instead of a large deviation event of probability decaying exponentially fast in tt. Let us begin by introducing a few notation. Let (Bt,t0)(B_{t},t\geq 0) be a standard Brownian motion, (τk,k1)(\tau_{k},k\geq 1) the atoms of an independent Poisson point process of intensity 22, and (Xu(k)(t),u𝒩t(k),t0)(X^{(k)}_{u}(t),u\in\mathcal{N}^{(k)}_{t},t\geq 0) i.i.d. BBMs, which are further independent of BB and τ\tau. For ϱ>2\varrho>\sqrt{2}, we set

𝒟~tϱ=δ0+k1𝟙{τk<t}u𝒩τk(k)δBτkϱτk+Xu(τk)and𝒟~ϱ=limt𝒟~tϱ.\widetilde{\mathcal{D}}_{t}^{\varrho}=\delta_{0}+\sum_{k\geq 1}\mathbbm{1}_{\left\{\tau_{k}<t\right\}}\sum_{u\in\mathcal{N}^{(k)}_{\tau_{k}}}\delta_{B_{\tau_{k}}-\varrho\tau_{k}+X_{u}(\tau_{k})}\quad\text{and}\quad\widetilde{\mathcal{D}}^{\varrho}=\lim_{t\to\infty}\widetilde{\mathcal{D}}_{t}^{\varrho}. (3.20)

In words, the process 𝒟~ϱ\widetilde{\mathcal{D}}^{\varrho} consists in making one particle starts from 0 and travels backwards in time according to a Brownian motion with drift ϱ\varrho. This particle gives birth to offspring at rate 22, each newborn child starting an independent BBM from its current position, forward in time. The point measure 𝒟~ϱ\widetilde{\mathcal{D}}^{\varrho} then consists of the position of all particles alive at time 0.

As a first step, we mention the following result, which can be thought of as a spinal decomposition argument with respect to the rightmost particle. This result can be found in [BBCM20, Lemma 2.1].

Proposition 3.5.

For all t0t\geq 0 set

t=u𝒩tδXu(t)Mt\mathcal{E}^{*}_{t}=\sum_{u\in\mathcal{N}_{t}}\delta_{X_{u}(t)-M_{t}}

the extremal process seen from the rightmost position. For all measurable non-negative functions f,Ff,F, we have

𝐄(f(Mtϱt)F(t))=e(1ϱ2/2)t𝐄(eϱBtf(Bt)F(𝒟~tϱ)𝟙{𝒟~tϱ((0,))=0})\mathbf{E}\left(f(M_{t}-\varrho t)F(\mathcal{E}^{*}_{t})\right)=e^{(1-\varrho^{2}/2)t}\mathbf{E}\left(e^{-\varrho B_{t}}f(B_{t})F(\widetilde{\mathcal{D}}_{t}^{\varrho})\mathbbm{1}_{\left\{\widetilde{\mathcal{D}}^{\varrho}_{t}((0,\infty))=0\right\}}\right)

It then follows from (3.16) and the above proposition that the law 𝒟ϱ\mathcal{D}^{\varrho} can be represented by conditioning the point measure 𝒟~ϱ\widetilde{\mathcal{D}}^{\varrho}, as was obtained in [BBCM20, Theorem 1.1].

Lemma 3.6.

For all ϱ>2\varrho>\sqrt{2},

  • the constant C(ϱ)C(\varrho) introduced in (3.16) is given by C(ϱ)=𝐏(𝒟~ϱ((0,))=0)C(\varrho)=\mathbf{P}(\widetilde{\mathcal{D}}^{\varrho}((0,\infty))=0).

  • the law of the point measure 𝒟ϱ\mathcal{D}^{\varrho} introduced in (3.19) can be constructed as

    𝐏(𝒟ϱ)=𝐏(𝒟~ϱ|𝒟~ϱ((0,))=0).\mathbf{P}(\mathcal{D}^{\varrho}\in\cdot)=\mathbf{P}(\widetilde{\mathcal{D}}^{\varrho}\in\cdot|\widetilde{\mathcal{D}}^{\varrho}((0,\infty))=0).

We end this section with an uniform estimate on the Laplace transform of the extremal process of the BBM, that generalizes both (3.16) and (3.19).

Lemma 3.7.

Let ϱ>2\varrho>\sqrt{2}, we set

tϱ(x)=u𝒩tδXu(t)ϱt+x.\mathcal{E}_{t}^{\varrho}(x)=\sum_{u\in\mathcal{N}_{t}}\delta_{X_{u}(t)-\varrho t+x}.

Let A>0A>0, for all φ𝒯\varphi\in\mathcal{T}, we have

𝐄(1etϱ(x),φ)=C(ϱ)e(1ϱ2/2)t2πteϱxx22teϱz(1eΨϱ[φ](z))dz(1+o(1)),\mathbf{E}\left(1-e^{-{\left\langle\mathcal{E}_{t}^{\varrho}(x),\varphi\right\rangle}}\right)=C(\varrho)\frac{e^{\left(1-\varrho^{2}/2\right)t}}{\sqrt{2\pi t}}e^{\varrho x-\tfrac{x^{2}}{2t}}\int e^{-\varrho z}\left(1-e^{-\Psi^{\varrho}[\varphi](z)}\right)\mathrm{d}z(1+o(1)),

uniformly in |x|At1/2|x|\leq At^{1/2}, as tt\to\infty, where Ψϱ[φ]:zlog𝐄(e𝒟ϱ,φ(+z))\Psi^{\varrho}[\varphi]:z\mapsto-\log\mathbf{E}\left(e^{-{\left\langle\mathcal{D}^{\varrho},\varphi(\cdot+z)\right\rangle}}\right).

Proof.

Let L>0L>0, recall from Lemma 3.2 that

𝐏(Mtϱtx+L)Ct1/2et(1ϱ2/2)eϱxx22teϱL.\mathbf{P}(M_{t}\geq\varrho t-x+L)\leq Ct^{-1/2}e^{t(1-\varrho^{2}/2)}e^{\varrho x-\tfrac{x^{2}}{2t}}e^{-\varrho L}.

Thus, as φ\varphi is non-negative, we have

0𝐄(1etϱ(x),φ)𝐄((1etϱ(x),φ)𝟙{Mtϱtx+L})Ct1/2et(1ϱ2/2)eϱxx22teϱL.0\leq\mathbf{E}\left(1-e^{-{\left\langle\mathcal{E}_{t}^{\varrho}(x),\varphi\right\rangle}}\right)-\mathbf{E}\left(\left(1-e^{-{\left\langle\mathcal{E}_{t}^{\varrho}(x),\varphi\right\rangle}}\right)\mathbbm{1}_{\left\{M_{t}\leq\varrho t-x+L\right\}}\right)\leq Ct^{-1/2}e^{t(1-\varrho^{2}/2)}e^{\varrho x-\tfrac{x^{2}}{2t}}e^{-\varrho L}. (3.21)

We also recall that the support of φ\varphi is bounded on the left, i.e. is included on [R,)[R,\infty) for some RR\in\mathbb{R}. Observe then that etϱ(x),φ=1e^{-{\left\langle\mathcal{E}_{t}^{\varrho}(x),\varphi\right\rangle}}=1 on the event {Mtϱtx+R}\{M_{t}\leq\varrho t-x+R\}.

We now use Proposition 3.5 to compute

𝐄((1etϱ(x),φ)𝟙{Mtϱt+x[R,L]})=et(1ϱ2/2)𝐄(eϱBt(1e𝒟~tϱ,τBt+xφ)𝟙{x+Bt[R,L],𝒟~tϱ((0,))=0}),\mathbf{E}\left(\left(1-e^{-{\left\langle\mathcal{E}_{t}^{\varrho}(x),\varphi\right\rangle}}\right)\mathbbm{1}_{\left\{M_{t}-\varrho t+x\in[R,L]\right\}}\right)\\ =e^{t(1-\varrho^{2}/2)}\mathbf{E}\left(e^{\varrho B_{t}}\left(1-e^{-{\left\langle\widetilde{\mathcal{D}}^{\varrho}_{t},\tau_{B_{t}+x}\varphi\right\rangle}}\right)\mathbbm{1}_{\left\{x+B_{t}\in[R,L],\widetilde{\mathcal{D}}^{\varrho}_{t}((0,\infty))=0\right\}}\right),

where τz(φ)()=φ(z+)\tau_{z}(\varphi)(\cdot)=\varphi(z+\cdot). Therefore, setting

Gt(x,z)=𝐄((1e𝒟~tϱ,τzφ)𝟙{𝒟~tϱ((0,))=0}|Bt=zx),G_{t}(x,z)=\mathbf{E}\left(\left(1-e^{-{\left\langle\widetilde{\mathcal{D}}^{\varrho}_{t},\tau_{z}\varphi\right\rangle}}\right)\mathbbm{1}_{\left\{\widetilde{\mathcal{D}}^{\varrho}_{t}((0,\infty))=0\right\}}\middle|B_{t}=z-x\right),

we have

et(ϱ2/21)2πteϱx+x22t𝐄((1etϱ(x),φ)𝟙{Mtϱt+x[R,L]})=RLeϱy+o(t1/2)Gt(x,y)dy,e^{t(\varrho^{2}/2-1)}\sqrt{2\pi t}e^{-\varrho x+\tfrac{x^{2}}{2t}}\mathbf{E}\left(\left(1-e^{-{\left\langle\mathcal{E}_{t}^{\varrho}(x),\varphi\right\rangle}}\right)\mathbbm{1}_{\left\{M_{t}-\varrho t+x\in[R,L]\right\}}\right)=\int_{R}^{L}e^{-\varrho y+o(t^{-1/2})}G_{t}(x,y)\mathrm{d}y, (3.22)

with the o(t1/2)o(t^{-1/2}) term being uniform in |x|At1/2|x|\leq At^{1/2}.

With the same computations as in the proof of [BBCM20, Lemma 3.4], we obtain

limtsup|x|At1/2Gt(x,y)=limtinf|x|At1/2Gt(x,y)\displaystyle\lim_{t\to\infty}\sup_{|x|\leq At^{1/2}}G_{t}(x,y)=\lim_{t\to\infty}\inf_{|x|\leq At^{1/2}}G_{t}(x,y) =𝐄((1e𝒟~ϱ,τyφ)𝟙{𝒟~ϱ((0,))=0})\displaystyle=\mathbf{E}\left(\left(1-e^{-{\left\langle\widetilde{\mathcal{D}}^{\varrho},\tau_{y}\varphi\right\rangle}}\right)\mathbbm{1}_{\left\{\widetilde{\mathcal{D}}^{\varrho}((0,\infty))=0\right\}}\right)
=C(ϱ)𝐄(1e𝒟ϱ,τyφ),\displaystyle=C(\varrho)\mathbf{E}\left(1-e^{-{\left\langle\mathcal{D}^{\varrho},\tau_{y}\varphi\right\rangle}}\right),

using the construction of 𝒟ϱ\mathcal{D}^{\varrho} given in Lemma 3.6. Therefore, using (3.21) and applying the dominated convergence theorem, equation (3.22) yields

lim suptsup|x|At1/2|et(ϱ2/21)2πteϱx+x22t𝐄((1etϱ(x),φ))C(ϱ)eϱy𝐄(1e𝒟ϱ,τyφ)dy|CeϱL,\limsup_{t\to\infty}\sup_{|x|\leq At^{1/2}}\left|e^{t(\varrho^{2}/2-1)}\sqrt{2\pi t}e^{-\varrho x+\tfrac{x^{2}}{2t}}\mathbf{E}\left(\left(1-e^{-{\left\langle\mathcal{E}_{t}^{\varrho}(x),\varphi\right\rangle}}\right)\right)-C(\varrho)\int_{\mathbb{R}}e^{-\varrho y}\mathbf{E}\left(1-e^{-{\left\langle\mathcal{D}^{\varrho},\tau_{y}\varphi\right\rangle}}\right)\mathrm{d}y\right|\\ \leq Ce^{-\varrho L},

which, letting LL\to\infty, completes the proof. ∎

Remark 3.8.

Note that applying Lemma 3.7 to function φ(z)=𝟙{z0}\varphi(z)=\mathbbm{1}_{\left\{z\geq 0\right\}} yields (3.16), and up simple computations, this lemma can also be used to obtain (3.19).

4 Multitype many-to-one lemmas

The many-to-one lemma is an ubiquitous result in the study of branching Brownian motions. This result links additive moments of the BBM with Brownian motion estimates. We first recall the classical version of this lemma, before giving a multitype version that applies to our process.

Let (Xu(t),u𝒩t)(X_{u}(t),u\in\mathcal{N}_{t}) be a standard BBM with branching rate 11. The classical many-to-one lemma can be tracked back at least to the work of Kahane and Peyrière [KP76, Pey74] on multiplicative cascades. It can be expressed as follows: for all t0t\geq 0 and measurable non-negative functions ff, we have

𝐄(u𝒩tf(Xu(s),st))=et𝐄(f(Bs,st)),\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}}f(X_{u}(s),s\leq t)\right)=e^{t}\mathbf{E}(f(B_{s},s\leq t)), (4.1)

where BB is a standard Brownian motion.

Recall that 𝒩t1\mathcal{N}^{1}_{t} (respectively 𝒩t2\mathcal{N}^{2}_{t}) is the set of particles of type 11 (resp. type 22) alive at time tt. Note that the process (Xu(t),u𝒩t1)t0(X_{u}(t),u\in\mathcal{N}^{1}_{t})_{t\geq 0} is a BBM with branching rate β\beta and diffusion σ2\sigma^{2}. Thus in view of (3.1), (4.1) implies that for all measurable non-negative function ff

𝐄(u𝒩t1f(Xu(s),st))=eβt𝐄(f(σBs,st)).\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}^{1}}f(X_{u}(s),s\leq t)\right)=e^{\beta t}\mathbf{E}(f(\sigma B_{s},s\leq t)).

Similarly, writing 𝐏(2)\mathbf{P}^{(2)} the law of the process starting from a single particle of type 22. As this particle behaves as in a standard BBM and only gives birth of particles of type 22, this process again is a BBM, therefore

𝐄(2)(u𝒩t2f(Xu(s),st))=et𝐄(f(Bs,st)),\mathbf{E}^{(2)}\left(\sum_{u\in\mathcal{N}_{t}^{2}}f(X_{u}(s),s\leq t)\right)=e^{t}\mathbf{E}(f(B_{s},s\leq t)),

writing 𝐄(2)\mathbf{E}^{(2)} for the expectation associated to 𝐏(2)\mathbf{P}^{(2)}.

The main aim of this section is to prove the following result, which allows to represent an additive functional of particles of type 22 appearing in the multitype BBM by a variable speed Brownian motion.

Proposition 4.1.

For all measurable non-negative function ff, we have

𝐄(u𝒩t2f((Xu(s),st),T(u)))=α0teβs+(ts)𝐄(f((σBus+(BuBus),ut),s))ds,\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}^{2}}f((X_{u}(s),s\leq t),T(u))\right)=\alpha\int_{0}^{t}e^{\beta s+(t-s)}\mathbf{E}\left(f((\sigma B_{u\wedge s}+(B_{u}-B_{u\wedge s}),u\leq t),s)\right)\mathrm{d}s,

where we recall that T(u)T(u) is the birth time of the first ancestor of type 22 of uu.

To prove this result, we begin by investigating the set \mathcal{B} of particles of type 22 that are born from a particle of type 11, that can be defined as

:={ut0𝒩2(t):T(u)=bu}.\mathcal{B}:=\left\{u\in\cup_{t\geq 0}\mathcal{N}^{2}(t):T(u)=b_{u}\right\}.

We observe that \mathcal{B} can be thought of as a Poisson point process with random intensity.

Lemma 4.2.

Conditionally on 1=σ(Xu(t),u𝒩t1,t0)\mathcal{F}^{1}=\sigma(X_{u}(t),u\in\mathcal{N}^{1}_{t},t\geq 0), the point measure uδ(Xu(s),sT(u))\displaystyle\sum_{u\in\mathcal{B}}\delta_{(X_{u}(s),s\leq T(u))} is a Poisson point process with intensity αdtu𝒩t1δ(Xu(s),st)\displaystyle\alpha\mathrm{d}t\otimes\sum_{u\in\mathcal{N}^{1}_{t}}\delta_{(X_{u}(s),s\leq t)}.

Proof.

This is a straightforward consequence of the definition of the two-type BBM and the superposition principle for Poisson process. Over its lifetime, a particle of type 11 gives birth to particles of type 22 according to a Poisson process with intensity α\alpha, and the trajectory leading to the newborn particle at time tt is exactly the same as the trajectory of its parent particle up to time tt. ∎

A direct consequence of the above lemma is the following applications of Poisson summation formula.

Corollary 4.3.

For all measurable non-negative function ff, we have

𝐄(uf(Xu(s),sT(u)))\displaystyle\mathbf{E}\left(\sum_{u\in\mathcal{B}}f(X_{u}(s),s\leq T(u))\right) =α0eβt𝐄(f(σBs,st))dt,\displaystyle=\alpha\int_{0}^{\infty}e^{\beta t}\mathbf{E}(f(\sigma B_{s},s\leq t))\mathrm{d}t, (4.2)
𝐄(exp(uf(Xu(s),sT(u))))\displaystyle\mathbf{E}\left(\exp\left(-\sum_{u\in\mathcal{B}}f(X_{u}(s),s\leq T(u))\right)\right) =𝐄(exp(α0u𝒩t11ef(Xu(s),st)dt))\displaystyle=\mathbf{E}\left(\exp\left(-\alpha\int_{0}^{\infty}\sum_{u\in\mathcal{N}_{t}^{1}}1-e^{-f(X_{u}(s),s\leq t)}\mathrm{d}t\right)\right) (4.3)
Proof.

We denote by 1=σ(Xu(s),u𝒩s1,s0)\mathcal{F}^{1}=\sigma(X_{u}(s),u\in\mathcal{N}^{1}_{s},s\geq 0) the filtration generated by all particles of type 11. We can compute

𝐄(uf(Xu(s),sT(u))|1)=α0u𝒩t1f(Xu(s),st)dt\mathbf{E}\left(\sum_{u\in\mathcal{B}}f(X_{u}(s),s\leq T(u))\middle|\mathcal{F}^{1}\right)=\alpha\int_{0}^{\infty}\sum_{u\in\mathcal{N}^{1}_{t}}f(X_{u}(s),s\leq t)\mathrm{d}t

using Lemma 4.2. Then using Fubini’s theorem and (4.1), we conclude that

𝐄(uf(Xu(s),sT(u)))=α0eβt𝐄(f(σBs,st))dt.\mathbf{E}\left(\sum_{u\in\mathcal{B}}f(X_{u}(s),s\leq T(u))\right)=\alpha\int_{0}^{\infty}e^{\beta t}\mathbf{E}(f(\sigma B_{s},s\leq t))\mathrm{d}t.

Similarly, using the exponential Poisson formula, we have

𝐄(exp(uf(Xu(s),sT(u)))|1)=exp(α0u𝒩t11ef(Xu(s),st)dt).\mathbf{E}\left(\exp\left(-\sum_{u\in\mathcal{B}}f(X_{u}(s),s\leq T(u))\right)\middle|\mathcal{F}^{1}\right)=\exp\left(-\alpha\int_{0}^{\infty}\sum_{u\in\mathcal{N}_{t}^{1}}1-e^{-f(X_{u}(s),s\leq t)}\mathrm{d}t\right).

Taking the expectation of this formula completes the proof of this corollary. ∎

We now turn to the proof of the multitype many-to-one lemma.

Proof of Proposition 4.1.

Let f,gf,g be two measurable bounded functions. For any u,uu,u^{\prime} particles in the BBM, we write uuu^{\prime}\succcurlyeq u to denote that uu^{\prime} is a descendant of uu. We compute

𝐄(u𝒩t2f(Xu(s),sT(u))g(Xu(s),s[T(u),t]))\displaystyle\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}^{2}}f(X_{u}(s),s\leq T(u))g(X_{u}(s),s\in[T(u),t])\right)
=\displaystyle= 𝐄(u𝟙{T(u)t}f(Xu(s),sT(u))u𝒩t2uug(Xu(s),s[T(u),t]))\displaystyle\mathbf{E}\left(\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{T(u)\leq t\right\}}f(X_{u}(s),s\leq T(u))\sum_{\begin{subarray}{c}u^{\prime}\in\mathcal{N}^{2}_{t}\\ u^{\prime}\succcurlyeq u\end{subarray}}g(X_{u^{\prime}}(s),s\in[T(u),t])\right)
=\displaystyle= 𝐄(u𝟙{T(u)t}f(Xu(s),sT(u))φ(T(u),Xu(T(u)))),\displaystyle\mathbf{E}\left(\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{T(u)\leq t\right\}}f(X_{u}(s),s\leq T(u))\varphi(T(u),X_{u}(T(u)))\right),

using the branching property for the BBM: every particle uu\in\mathcal{B} starts an independent BBM from time T(u)T(u) and position Xu(T(u))X_{u}(T(u)). Here, we have set for xx\in\mathbb{R} and s0s\geq 0

φ(s,x)\displaystyle\varphi(s,x) =𝐄(2)(u𝒩ts2g(x+Xu(rs),r[s,t]))\displaystyle=\mathbf{E}^{(2)}\left(\sum_{u\in\mathcal{N}^{2}_{t-s}}g\left(x+X_{u}(r-s),r\in[s,t]\right)\right)
=ets𝐄(g(x+Brs,r[s,t])),\displaystyle=e^{t-s}\mathbf{E}\left(g\left(x+B_{r-s},r\in[s,t]\right)\right),

by the standard many-to-one lemma. Additionally, by Corollary 4.3, we have

𝐄(u𝒩t2f(Xu(s),sT(u))g(Xu(s),s[T(u),t]))\displaystyle\mathbf{E}\left(\sum_{u\in\mathcal{N}_{t}^{2}}f(X_{u}(s),s\leq T(u))g(X_{u}(s),s\in[T(u),t])\right)
=\displaystyle= α0teβs𝐄(f(σBr,rs)φ(s,σBs))ds\displaystyle\alpha\int_{0}^{t}e^{\beta s}\mathbf{E}(f(\sigma B_{r},r\leq s)\varphi(s,\sigma B_{s}))\mathrm{d}s
=\displaystyle= α0teβs+ts𝐄(f(σBr,rs)g(σBs+(BrBs),r[s,t]))ds.\displaystyle\alpha\int_{0}^{t}e^{\beta s+t-s}\mathbf{E}(f(\sigma B_{r},r\leq s)g\left(\sigma B_{s}+(B_{r}-B_{s}),r\in[s,t])\right)\mathrm{d}s.

Using the monotone class theorem, the proof of Proposition 4.1 is now complete. ∎

5 Proof of Theorem 1.2

We assume in this section that (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, i.e. that either σ2>1\sigma^{2}>1 and σ2<1β\sigma^{2}<\frac{1}{\beta} or σ21\sigma^{2}\leq 1 and σ2<2β\sigma^{2}<2-\beta. In that case, we show that the extremal process is dominated by the behaviour of particles of type 22 that are born at the beginning of the process. The main steps of the proof of Theorem 1.2 are the following:

  1. 1.

    We show that for all A>0A>0, there exists R>0R>0 such that with high probability, every particle uu of type 22 to the right of mt(II)Am^{(II)}_{t}-A satisfy T(u)RT(u)\leq R.

  2. 2.

    We use the convergence in distribution of the extremal process of a single-type branching Brownian motion to demonstrate that the extremal process generated by the individuals born of type 22 before time RR converges as tt\to\infty.

  3. 3.

    We prove that letting RR\to\infty, the above extremal process converges, and the limiting point measure is the point measure of the full two-type branching Brownian motion.

In this section, we write v=2βσ2v=\sqrt{2\beta\sigma^{2}} and θ=2β/σ2\theta=\sqrt{2\beta/\sigma^{2}}, which are respectively the speed and critical parameter of the branching Brownian motion of particles of type 11. Recall that mt(II)=2t322logtm_{t}^{(II)}=\sqrt{2}t-\frac{3}{2\sqrt{2}}\log t, we write

^t=u𝒩t2δXu(t)mt(II),\widehat{\mathcal{E}}_{t}=\sum_{u\in\mathcal{N}^{2}_{t}}\delta_{X_{u}(t)-m^{(II)}_{t}},

the extremal process of particles of type 22 in the branching Brownian motion, centred around mt(II)m^{(II)}_{t}. We begin by proving that with high probability, no particle of type 22 that was born from a particle of type 11 after time RR has a descendant close to mt(II)m_{t}^{(II)}.

Lemma 5.1.

Assuming that (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II}, for all A>0A>0, we have

limRlim supt𝐏(u𝒩t2:T(u)R,Xu(t)mt(II)A)=0.\lim_{R\to\infty}\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:T(u)\geq R,X_{u}(t)\geq m_{t}^{(II)}-A)=0.
Proof.

Let K>0K>0, we first recall that by (3.1) and (3.13), we have

𝐏(t0,u𝒩t1:Xu(t)vt+K)eθK,\mathbf{P}\left(\exists t\geq 0,u\in\mathcal{N}^{1}_{t}:X_{u}(t)\geq vt+K\right)\leq e^{-\theta K},

i.e. that with high probability, all particles of type 11 stay below the curve svs+Ks\mapsto vs+K.

We now set, for R,A,K0R,A,K\geq 0 and t0t\geq 0:

Yt(A,R,K)=u𝟙{T(u)>R,Xu(T(u))vT(u)+K}𝟙{Mtumt(II)A},Y_{t}(A,R,K)=\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{T(u)>R,X_{u}(T(u))\leq vT(u)+K\right\}}\mathbbm{1}_{\left\{M^{u}_{t}\geq m_{t}^{(II)}-A\right\}},

where MtuM^{u}_{t} is the position of the rightmost descendant at time tt of the individual uu. In other words, Yt(A,R,K)Y_{t}(A,R,K) is the number of particles of type 22 born from a particle of type 11 after time RR, that were born below the curve svs+Ks\mapsto vs+K and have a member of their family to the right of mt(II)Am_{t}^{(II)}-A. Observe that by Markov inequality, we have

𝐏(u𝒩t2:T(u)R,Xu(t)mt(II)A)\displaystyle\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:T(u)\geq R,X_{u}(t)\geq m_{t}^{(II)}-A) 𝐏(t0,u𝒩t1:Xu(t)vs+K)+𝐏(Yt(A,R,K)1)\displaystyle\leq\mathbf{P}\left(\exists t\geq 0,u\in\mathcal{N}^{1}_{t}:X_{u}(t)\geq vs+K\right)+\mathbf{P}(Y_{t}(A,R,K)\geq 1)
eθK+𝐄(Yt(A,R,K)).\displaystyle\leq e^{-\theta K}+\mathbf{E}(Y_{t}(A,R,K)).

To complete the proof, it is therefore enough to bound lim supt𝐄(Yt(A,R,K))\limsup_{t\to\infty}\mathbf{E}(Y_{t}(A,R,K)). Using the branching property and Corollary 4.3, we have

𝐄(Yt(A,R,K))\displaystyle\mathbf{E}(Y_{t}(A,R,K)) =𝐄(u𝟙{T(u)[R,t]}𝟙{Xu(T(u))vT(u)+K}F(tT(u),Xu(T(u))))\displaystyle=\mathbf{E}\left(\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{T(u)\in[R,t]\right\}}\mathbbm{1}_{\left\{X_{u}(T(u))\leq vT(u)+K\right\}}F\left(t-T(u),X_{u}(T(u))\right)\right)
=αRteβs𝐄(F(ts,σBs)𝟙{σBsvs+K})ds,\displaystyle=\alpha\int_{R}^{t}e^{\beta s}\mathbf{E}\left(F\left(t-s,\sigma B_{s}\right)\mathbbm{1}_{\left\{\sigma B_{s}\leq vs+K\right\}}\right)\mathrm{d}s, (5.1)

where we have set F(r,x)=𝐏(2)(x+Mrmt(II)A)F(r,x)=\mathbf{P}^{(2)}\left(x+M_{r}\geq m_{t}^{(II)}-A\right).

By (3.15), there exists C>0C>0 such that for all xx\in\mathbb{R} and t0t\geq 0, we have

𝐏(2)(Mtmt(II)+x)C(1+x+)e2x,\mathbf{P}^{(2)}\left(M_{t}\geq m_{t}^{(II)}+x\right)\leq C(1+x_{+})e^{-\sqrt{2}x},

so that for all sts\leq t,

F(ts,x)\displaystyle F(t-s,x) =𝐏(2)(Mtsmts(II)+2s+322logts+1t+1Ax)\displaystyle=\mathbf{P}^{(2)}\left(M_{t-s}\geq m_{t-s}^{(II)}+\sqrt{2}s+\tfrac{3}{2\sqrt{2}}\log\tfrac{t-s+1}{t+1}-A-x\right)
C(t+1ts+1)32(1+2s+(x)+)e2(2sxA).\displaystyle\leq C\left(\frac{t+1}{t-s+1}\right)^{\frac{3}{2}}\left(1+\sqrt{2}s+(-x)_{+}\right)e^{-\sqrt{2}(\sqrt{2}s-x-A)}. (5.2)

We bound 𝐄(Yt(A,R,K))\mathbf{E}(Y_{t}(A,R,K)) in two different ways, depending on the sign of σ21\sigma^{2}-1.

First, if σ21\sigma^{2}\leq 1, we observe that the condition Xu(s)vs+KX_{u}(s)\leq vs+K does not play a major role in the asymptotic behaviour of 𝐄(Yt(A,R,K))\mathbf{E}(Y_{t}(A,R,K)). As a result, (5.1) and (5.2) yield

𝐄(Yt(A,R,K))CeARt(t+1ts+1)3/2es(β2)𝐄((1+2s+σ(Bs)+)e2σBs)ds,\mathbf{E}(Y_{t}(A,R,K))\leq Ce^{A}\int_{R}^{t}\left(\frac{t+1}{t-s+1}\right)^{3/2}e^{s(\beta-2)}\mathbf{E}\left(\left(1+\sqrt{2}s+\sigma(-B_{s})_{+}\right)e^{\sqrt{2}\sigma B_{s}}\right)\mathrm{d}s,

and as 𝐄((1+2s+σ(Bs)+)e2σBs)C(1+s)eσ2s\mathbf{E}\left(\left(1+\sqrt{2}s+\sigma(-B_{s})_{+}\right)e^{\sqrt{2}\sigma B_{s}}\right)\leq C(1+s)e^{\sigma^{2}s}, we have

𝐄(Yt(A,R,K))CeARt(t+1ts+1)32(s+1)exp(s(β+σ22))ds.\mathbf{E}(Y_{t}(A,R,K))\leq Ce^{A}\int_{R}^{t}\left(\frac{t+1}{t-s+1}\right)^{\frac{3}{2}}(s+1)\exp\left(s\left(\beta+\sigma^{2}-2\right)\right)\mathrm{d}s.

Hence, as (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II} and σ21\sigma^{2}\leq 1, we have β+σ22<0\beta+\sigma^{2}-2<0. Therefore, by dominated convergence theorem,

lim supt𝐄(Yt(A,R,K))CeAR(s+1)exp(s(β+σ22))ds,\limsup_{t\to\infty}\mathbf{E}(Y_{t}(A,R,K))\leq Ce^{A}\int_{R}^{\infty}(s+1)\exp\left(s\left(\beta+\sigma^{2}-2\right)\right)\mathrm{d}s,

which goes to 0 as RR\to\infty, completing the proof in that case.

We now assume that σ2>1\sigma^{2}>1. In that case, the condition Xu(s)2βσ2s+KX_{u}(s)\leq\sqrt{2\beta\sigma^{2}}s+K is needed to keep our upper bound small enough, as events of the form {Xu(s)vs}\{X_{u}(s)\geq vs\} have small probability but Yt(A,R,K)Y_{t}(A,R,K) is large on that event. Using the Girsanov transform, (5.1) yields

𝐄(Yt(A,R,K))\displaystyle\mathbf{E}(Y_{t}(A,R,K))
\displaystyle\leq αRt𝐄(eθσBsF(ts,σBs+vs)𝟙{σBsK})ds\displaystyle\alpha\int_{R}^{t}\mathbf{E}\left(e^{-\theta\sigma B_{s}}F(t-s,\sigma B_{s}+vs)\mathbbm{1}_{\left\{\sigma B_{s}\leq K\right\}}\right)\mathrm{d}s
\displaystyle\leq Cαe2ARte2(2v)s(t+1ts+1)32𝐄(e(2θ)σBs(1+(v+2)s+(Bs)+)𝟙{BsK})ds,\displaystyle C\alpha e^{\sqrt{2}A}\int_{R}^{t}e^{-\sqrt{2}(\sqrt{2}-v)s}\left(\frac{t+1}{t-s+1}\right)^{\frac{3}{2}}\mathbf{E}\left(e^{(\sqrt{2}-\theta)\sigma B_{s}}\left(1+(v+\sqrt{2})s+(-B_{s})_{+}\right)\mathbbm{1}_{\left\{B_{s}\leq K\right\}}\right)\mathrm{d}s,

using (5.2). As (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II} and σ2>1\sigma^{2}>1, we have βσ2<1\beta\sigma^{2}<1. This yields in particular β<σ2\beta<\sigma^{2} hence 2θ>0\sqrt{2}-\theta>0. Integrating with respect to the Brownian density, there exists C>0C>0 such that

𝐄(e(2θ)σBs(1+2(βσ2+1)s+(Bs)+)𝟙{BsK})C(1+s)12e(2θ)σK,\mathbf{E}\left(e^{(\sqrt{2}-\theta)\sigma B_{s}}\left(1+\sqrt{2}(\sqrt{\beta\sigma^{2}}+1)s+(-B_{s})_{+}\right)\mathbbm{1}_{\left\{B_{s}\leq K\right\}}\right)\leq C(1+s)^{\frac{1}{2}}e^{(\sqrt{2}-\theta)\sigma K},

yielding

𝐄(Yt(A,R,K))Cαe2A+(2θ)σKRte2(1βσ2)s(t+1ts+1)32(1+s)12ds.\mathbf{E}(Y_{t}(A,R,K))\leq C\alpha e^{\sqrt{2}A+(\sqrt{2}-\theta)\sigma K}\int_{R}^{t}e^{-2(1-\sqrt{\beta\sigma^{2}})s}\left(\frac{t+1}{t-s+1}\right)^{\frac{3}{2}}(1+s)^{\frac{1}{2}}\mathrm{d}s.

Then by dominated convergence, as 1βσ2>01-\sqrt{\beta\sigma^{2}}>0, we deduce that

lim supt𝐄(Yt(A,R,K))Cαe2A+(2θ)σKRe2(1βσ2)s(1+s)12ds,\limsup_{t\to\infty}\mathbf{E}(Y_{t}(A,R,K))\leq C\alpha e^{\sqrt{2}A+(\sqrt{2}-\theta)\sigma K}\int_{R}^{\infty}e^{-2(1-\sqrt{\beta\sigma^{2}})s}(1+s)^{\frac{1}{2}}\mathrm{d}s,

which decreases to 0 as RR\to\infty, completing the proof. ∎

We now use the known asymptotic behaviour of the extremal process of the branching Brownian motion, recalled in Section 3, to compute the asymptotic behaviour of the extremal process of particles satisfying T(u)RT(u)\leq R, defined as

^tR:=u𝒩t2𝟙{T(u)R}δXu(t)mt(II).\widehat{\mathcal{E}}^{R}_{t}:=\sum_{u\in\mathcal{N}^{2}_{t}}\mathbbm{1}_{\left\{T(u)\leq R\right\}}\delta_{X_{u}(t)-m^{(II)}_{t}}.

For any uu\in\mathcal{B}, and t0t\geq 0, we set

Zt(u):=u𝒩t2uu(2tXu(t))e2(Xu(t)2t),Z^{(u)}_{t}:=\sum_{\begin{subarray}{c}u^{\prime}\in\mathcal{N}^{2}_{t}\\ u^{\prime}\succcurlyeq u\end{subarray}}(\sqrt{2}t-X_{u^{\prime}}(t))e^{\sqrt{2}(X_{u^{\prime}}(t)-\sqrt{2}t)},

where we recall that uuu^{\prime}\succcurlyeq u denotes that uu^{\prime} is a descendant of uu. Note that by (3.9) and the branching property, Zt(u)Z^{(u)}_{t} converges a.s. to the variable Z(u):=lim inftZt(u).\displaystyle Z^{(u)}_{\infty}:=\liminf_{t\to\infty}Z^{(u)}_{t}. Moreover, e2(Xu(T(u))2T(u))Z(u)=(d)Ze^{-\sqrt{2}(X_{u}(T(u))-\sqrt{2}T(u))}Z^{(u)}_{\infty}{\overset{(d)}{=}}Z_{\infty}, where ZZ_{\infty} is the limit of the derivative martingale of a standard branching Brownian motion.

Lemma 5.2.

For all φ𝒯\varphi\in\mathcal{T}, we have limt^tR,φ=^R,φ\displaystyle\lim_{t\to\infty}{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}={\left\langle\widehat{\mathcal{E}}^{R}_{\infty},\varphi\right\rangle} in law, where ^R\widehat{\mathcal{E}}^{R}_{\infty} is a decorated Poisson point process with intensity cZ¯R2e2xdxc_{\star}\overline{Z}_{R}\sqrt{2}e^{-\sqrt{2}x}\mathrm{d}x and decoration law 𝔇\mathfrak{D}, with Z¯R:=u𝟙{T(u)R}Z(u)\displaystyle\overline{Z}_{R}:=\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{T(u)\leq R\right\}}Z^{(u)}_{\infty}.

Proof.

Let φ𝒯\varphi\in\mathcal{T} be a test function. Observe that using the branching property of the branching Brownian motion, we have

𝐄(exp(^tR,φ))=𝐄(u:T(u)RFt(T(u),Xu(T(u)))),\mathbf{E}\left(\exp\left(-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}\right)\right)=\mathbf{E}\left(\prod_{u\in\mathcal{B}:T(u)\leq R}F_{t}(T(u),X_{u}(T(u)))\right),

where Ft(s,x)=𝐄(2)(exp(u𝒩tφ(x+Xu(ts)mt(II))))F_{t}(s,x)=\mathbf{E}^{(2)}\left(\exp\left(-\sum_{u\in\mathcal{N}_{t}}\varphi\left(x+X_{u}(t-s)-m_{t}^{(II)}\right)\right)\right) for 0st0\leq s\leq t and xx\in\mathbb{R}. Using again that mt(II)=mts(II)+2s+o(1)m_{t}^{(II)}=m_{t-s}^{(II)}+\sqrt{2}s+o(1) as tt\to\infty and applying Lemma 3.4, we have for all s0s\geq 0

limtFt(s,x)=𝐄(2)(exp(cZe2x2s(1eΨ[φ](z))2e2zdz)),\lim_{t\to\infty}F_{t}(s,x)=\mathbf{E}^{(2)}\left(\exp\left(-c_{\star}Z_{\infty}e^{\sqrt{2}x-2s}\int(1-e^{-\Psi[\varphi](z)})\sqrt{2}e^{-\sqrt{2}z}\mathrm{d}z\right)\right),

where ZZ_{\infty} is the limit of the derivative martingale in a standard branching Brownian motion. Therefore, by dominated convergence theorem,

limt𝐄(exp(^tR,φ))=𝐄(exp(cZ¯R(1eΨ[φ](z))2e2zdz)),\lim_{t\to\infty}\mathbf{E}\left(\exp\left(-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}\right)\right)=\mathbf{E}\left(\exp\left(-c_{\star}\overline{Z}_{R}\int(1-e^{-\Psi[\varphi](z)})\sqrt{2}e^{-\sqrt{2}z}\mathrm{d}z\right)\right),

with Z¯R=u𝟙{T(u)R}Z(u)\overline{Z}_{R}=\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{T(u)\leq R\right\}}Z_{\infty}^{(u)}, completing the proof. ∎

We then observe that ^R\widehat{\mathcal{E}}^{R}_{\infty} converges in law as RR\to\infty to (II)\mathcal{E}^{(II)}_{\infty} the point measure defined in Theorem 1.2.

Lemma 5.3.

For all φ𝒯\varphi\in\mathcal{T}, we have limR^R,φ=(II),φ\lim_{R\to\infty}{\left\langle\widehat{\mathcal{E}}^{R}_{\infty},\varphi\right\rangle}={\left\langle\mathcal{E}^{(II)}_{\infty},\varphi\right\rangle} in law, where Z¯:=uZ(u)\overline{Z}_{\infty}:=\sum_{u\in\mathcal{B}}Z^{(u)}_{\infty}.

Proof.

Recall that Z0Z_{\infty}\geq 0 a.s. therefore (Z¯R,R0)(\overline{Z}_{R},R\geq 0) is increasing and Z¯=limRZ¯R\overline{Z}_{\infty}=\lim_{R\to\infty}\overline{Z}_{R} exists a.s. Given that for all function φ𝒯\varphi\in\mathcal{T},

𝐄(e^R,φ)=𝐄(exp(cZ¯R(1eΨ[φ](z))2e2zdz)),\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{\infty},\varphi\right\rangle}}\right)=\mathbf{E}\left(\exp\left(-c_{\star}\overline{Z}_{R}\int(1-e^{-\Psi[\varphi](z)})\sqrt{2}e^{-\sqrt{2}z}\mathrm{d}z\right)\right),

to prove that R\mathcal{E}^{R}_{\infty} converges in law, it is enough to show that Z¯<\overline{Z}_{\infty}<\infty a.s.

We recall that v=2βσ2v=\sqrt{2\beta\sigma^{2}} the speed of the branching Brownian motion of particles of type 11. As

limtmaxu𝒩t1Xu(t)vt=a.s.\lim_{t\to\infty}\max_{u\in\mathcal{N}^{1}_{t}}X_{u}(t)-vt=-\infty\quad\text{a.s.}

(which is a consequence of the fact that the additive martingale at the critical parameter converges to 0 a.s.), there is almost surely finitely many uu\in\mathcal{B} with Xu(T(u))vT(u)X_{u}(T(u))\geq vT(u). To prove the finiteness of Z¯\overline{Z}_{\infty}, we then use the following variation on Kolmogorov’s three series theorem: If we have

u𝟙{Xu(T(u))vT(u)}𝐄(Z(u)1|1σ())<a.s.,\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{X_{u}(T(u))\leq vT(u)\right\}}\mathbf{E}\left(Z_{\infty}^{(u)}\wedge 1\middle|\mathcal{F}^{1}\vee\sigma(\mathcal{B})\right)<\infty\quad\text{a.s.}, (5.3)

then Z~:=u𝟙{Xu(T(u))vT(u)}Z(u)<\widetilde{Z}_{\infty}:=\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{X_{u}(T(u))\leq vT(u)\right\}}Z_{\infty}^{(u)}<\infty a.s, where we recall that 1=σ(Xu(t),u𝒩t,t0)\mathcal{F}^{1}=\sigma(X_{u}(t),u\in\mathcal{N}_{t},t\geq 0). Using that Z¯\overline{Z}_{\infty} is obtained by adding a finite number of finite random variables to Z~\widetilde{Z}_{\infty}, it implies that Z¯<\overline{Z}_{\infty}<\infty a.s

Indeed, if we assume (5.3), using the Markov inequality and the Borel-Catelli lemma, we deduce that almost surely there are finitely many uu\in\mathcal{B} whose contribution to Z~\widetilde{Z}_{\infty} is larger than 11. Additionally, (5.3) also implies that the sum of all the other contributions to Z~\widetilde{Z}_{\infty} has finite mean. Hence, we have Z~<\widetilde{Z}_{\infty}<\infty a.s.

We now prove (5.3), using that by (3.10) and (3.11) for all xx\in\mathbb{R}, we have 𝐄((Zex)1)C(1+(x)+)ex\mathbf{E}((Z_{\infty}e^{x})\wedge 1)\leq C(1+(-x)_{+})e^{x}. Hence, using that Z(u)=(d)e2(Xu(T(u))2T(u))ZZ^{(u)}_{\infty}{\overset{(d)}{=}}e^{\sqrt{2}(X_{u}(T(u))-\sqrt{2}T(u))}Z_{\infty} it is enough to show that

u𝟙{Xu(T(u))vT(u)}(1+(2T(u)2Xu(T(u)))+)e2Xu(T(u))2T(u)<a.s.\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{X_{u}(T(u))\leq vT(u)\right\}}\left(1+\left(2T(u)-\sqrt{2}X_{u}(T(u))\right)_{+}\right)e^{\sqrt{2}X_{u}(T(u))-2{T(u)}}<\infty\quad\text{a.s.} (5.4)

This quantity being a series of positive random variables, we prove that this series has finite mean to conclude. By Corollary 4.3, we have

𝐄(u𝟙{Xu(T(u))vT(u)}(1+(2T(u)2Xu(T(u)))+)e2Xu(T(u))2T(u))=α0eβs𝐄(𝟙{σBsvs}(1+(2s2σBs)+)e2σBs2s)ds.\mathbf{E}\left(\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{X_{u}(T(u))\leq v{T(u)}\right\}}\left(1+\left(2{T(u)}-\sqrt{2}X_{u}(T(u))\right)_{+}\right)e^{\sqrt{2}X_{u}(T(u))-2{T(u)}}\right)\\ =\alpha\int_{0}^{\infty}e^{\beta s}\mathbf{E}\left(\mathbbm{1}_{\left\{\sigma B_{s}\leq vs\right\}}\left(1+\left(2s-\sqrt{2}\sigma B_{s}\right)_{+}\right)e^{\sqrt{2}\sigma B_{s}-2s}\right)\mathrm{d}s.

Similarly to the proof of Lemma 5.1, we bound the above quantity in two different ways depending on whether σ2>1\sigma^{2}>1 or σ21\sigma^{2}\leq 1.

If σ21\sigma^{2}\leq 1, we have

eβs𝐄((1+(2σBs2s)+)e2σBs2s)C(1+s)exp(s(σ2+β2)),e^{\beta s}\mathbf{E}\left(\left(1+\left(\sqrt{2}\sigma B_{s}-2s\right)_{+}\right)e^{\sqrt{2}\sigma B_{s}-2s}\right)\leq C(1+s)\exp\left(s\left(\sigma^{2}+\beta-2\right)\right),

which decays exponentially fast as (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II} and σ21\sigma^{2}\leq 1. Therefore, we have

𝐄(u𝟙{Xu(T(u))vT(u)}(1+(2T(u)2Xu(T(u)))+)e2Xu(T(u))2T(u))<,\mathbf{E}\left(\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{X_{u}(T(u))\leq v{T(u)}\right\}}\left(1+\left(2T(u)-\sqrt{2}X_{u}(T(u))\right)_{+}\right)e^{\sqrt{2}X_{u}(T(u))-2{T(u)}}\right)<\infty,

proving (5.4), hence (5.3), therefore that Z¯<\overline{Z}_{\infty}<\infty a.s. in that case.

If σ2>1\sigma^{2}>1, we have

eβs𝐄(𝟙{σBsvs}(1+(2s2σBs)+)e2σBs2s)=𝐄(𝟙{Bs0}(1+(2(1σ2β)s2σBs)+)e2σBs)e2(σ2β1)sC(1+s)exp(s(σ2β1)).e^{\beta s}\mathbf{E}\left(\mathbbm{1}_{\left\{\sigma B_{s}\leq vs\right\}}\left(1+\left(2s-\sqrt{2}\sigma B_{s}\right)_{+}\right)e^{\sqrt{2}\sigma B_{s}-2s}\right)\\ =\mathbf{E}\left(\mathbbm{1}_{\left\{B_{s}\leq 0\right\}}\left(1+\left(2(1-\sqrt{\sigma^{2}\beta})s-\sqrt{2}\sigma B_{s}\right)_{+}\right)e^{\sqrt{2}\sigma B_{s}}\right)e^{2\left(\sqrt{\sigma^{2}\beta}-1\right)s}\\ \leq C(1+s)\exp\left(s\left(\sqrt{\sigma^{2}\beta}-1\right)\right).

As (β,σ2)𝒞II(\beta,\sigma^{2})\in\mathcal{C}_{II} and σ2<1\sigma^{2}<1 we have once again

𝐄(u𝟙{Xu(T(u))vT(u)}(1+(2T(u)2Xu(T(u)))+)e2Xu(T(u))2T(u))<,\mathbf{E}\left(\sum_{u\in\mathcal{B}}\mathbbm{1}_{\left\{X_{u}(T(u))\leq v{T(u)}\right\}}\left(1+\left(2{T(u)}-\sqrt{2}X_{u}(T(u))\right)_{+}\right)e^{\sqrt{2}X_{u}(T(u))-2{T(u)}}\right)<\infty,

which proves that Z¯<\overline{Z}_{\infty}<\infty a.s. in that case as well. ∎

Using the above results, we finally obtain the asymptotic behaviour of the extremal process in case 𝒞II\mathcal{C}_{II}.

Proof of Theorem 1.2.

Recall that ^t=u𝒩t2δXu(t)mt\widehat{\mathcal{E}}_{t}=\sum_{u\in\mathcal{N}^{2}_{t}}\delta_{X_{u}(t)-m_{t}}. Using Proposition 3.3, we only need to prove that for all φ𝒯\varphi\in\mathcal{T}, we have

limt𝐄(e^t,φ)=𝐄(e(II),φ).\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(e^{-{\left\langle\mathcal{E}^{(II)}_{\infty},\varphi\right\rangle}}\right).

Let φ𝒯\varphi\in\mathcal{T}, and set AA\in\mathbb{R} such that φ(z)=0\varphi(z)=0 for all zAz\leq A. By Lemma 5.1, for all ε>0\varepsilon>0, there exists R0R\geq 0 such that 𝐏(^tR(φ)^t(φ))ε.\mathbf{P}\left(\widehat{\mathcal{E}}^{R}_{t}(\varphi)\neq\widehat{\mathcal{E}}_{t}(\varphi)\right)\leq\varepsilon. Then, using that φ\varphi is non-negative, so that ^tR,φ^t,φ{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}\leq{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle} and e^t,φe^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}} is bounded by 11, we have 𝐄(e^tR,φ)𝐄(e^t,φ)𝐄(e^tR,φ)+ε\displaystyle\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)\leq\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)\leq\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)+\varepsilon. Applying Lemma 5.2 and Lemma 5.3 to let tt, then RR, grow to \infty, we obtain

𝐄(e(II),φ)lim inft𝐄(e^t,φ)lim supt𝐄(e^t,φ)𝐄(e(II),φ)+ε.\mathbf{E}\left(e^{-{\left\langle\mathcal{E}^{(II)}_{\infty},\varphi\right\rangle}}\right)\leq\liminf_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)\leq\limsup_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)\leq\mathbf{E}\left(e^{-{\left\langle\mathcal{E}^{(II)}_{\infty},\varphi\right\rangle}}\right)+\varepsilon.

Letting ε0\varepsilon\to 0 we obtain that limt𝐄(e^t,φ)=𝐄(e(II),φ)\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(e^{-{\left\langle\mathcal{E}^{(II)}_{\infty},\varphi\right\rangle}}\right) for all φ𝒯\varphi\in\mathcal{T}, which completes the proof of Theorem 1.2 by Remark 1.5. ∎

We end this section by conjecturing a possible direct formula for the computation of Z¯\overline{Z}_{\infty} as the limit of a sub-martingale of the multitype BBM.

Conjecture 5.4.

We have limtu𝒩t2(2tXu(t))e2Xu(t)2t=Z¯\lim_{t\to\infty}\sum_{u\in\mathcal{N}_{t}^{2}}(\sqrt{2}t-X_{u}(t))e^{\sqrt{2}X_{u}(t)-2t}=\overline{Z}_{\infty} a.s.

6 Proof of Theorem 1.1

In this section, we assume that (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, that is either σ21\sigma^{2}\leq 1 and σ2>1β\sigma^{2}>\frac{1}{\beta}, or σ2>1\sigma^{2}>1 and σ2>β2β1\sigma^{2}>\frac{\beta}{2\beta-1}. In that situation, we show that the extremal process of particles of type 22 is mainly driven by the asymptotic of particles of type 11, and that any particle of type 22 significantly contributing to the extremal process at time tt satisfies tT(u)=O(1)t-T(u)=O(1), meaning that they have a close ancestor of type 11.

For the rest of the section, we denote by v=2βσ2v=\sqrt{2\beta\sigma^{2}} and θ=2β/σ2\theta=\sqrt{2\beta/\sigma^{2}} the speed and critical parameter of the BBM of particles of type 11. Recall that mt(I)=vt32θlogtm^{(I)}_{t}=vt-\frac{3}{2\theta}\log t, and we set

^t:=u𝒩t2δXu(t)mt(I),\widehat{\mathcal{E}}_{t}:=\sum_{u\in\mathcal{N}^{2}_{t}}\delta_{X_{u}(t)-m^{(I)}_{t}},

the extremal process of particles of type 22, centred around mt(I)m^{(I)}_{t}.

To prove Theorem 1.1, we first show that for all φ𝒯\varphi\in\mathcal{T}, ^t,φ{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle} converges, as tt\to\infty to a proper random variable. By [Kal02, Lemma 5.1], this is enough to conclude that ^t\widehat{\mathcal{E}}_{t} converges vaguely in law to a limiting point measure ¯\overline{\mathcal{E}}. We then use that with high probability, no particle of type 22 born before time RR contributes, to the extremal process of the multitype BBM. Then, by the branching property, it shows that ¯\overline{\mathcal{E}} satisfies a stability under superposition probability which, by [Mai13, Corollary 3.2], can be identified as a decorated Poisson point process with intensity proportional to Z(I)eθxdxZ^{(I)}_{\infty}e^{-\theta x}\mathrm{d}x.

To prove the results of this section, we make use of the following extension of (3.13). For all t0t\geq 0, we write at=32θlog(t+1)a_{t}=\frac{3}{2\theta}\log(t+1). There exists C>0C>0 such that for all t0t\geq 0 and K>0K>0, we have

𝐏(st,u𝒩s1:Xu(s)vsat+ats+K)C(K+1)eθK.\mathbf{P}\left(\exists s\leq t,u\in\mathcal{N}^{1}_{s}:X_{u}(s)\geq vs-a_{t}+a_{t-s}+K\right)\leq C(K+1)e^{-\theta K}. (6.1)

This result was proved in [Mal15a] in the context of branching random walks, and has been adapted to continuous-time settings in [Mal15c, Lemma 3.1].

We first show the tightness of the law of the number of particles of type 22 born to the right of mt(I)Am^{(I)}_{t}-A.

Lemma 6.1.

We assume that (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}. For all A,K>0A,K>0, there exists CA,K>0C_{A,K}>0 and δ>0\delta>0 such that for all R0R\geq 0, we have

lim supt𝐄(u𝒩t2𝟙{Xu(t)mt(I)A}𝟙{T(u)tR}𝟙{Xu(s)vsat+ats+K,sT(u)})CA,KeδR.\limsup_{t\to\infty}\mathbf{E}\left(\sum_{u\in\mathcal{N}^{2}_{t}}\mathbbm{1}_{\left\{X_{u}(t)\geq m^{(I)}_{t}-A\right\}}\mathbbm{1}_{\left\{T(u)\leq t-R\right\}}\mathbbm{1}_{\left\{X_{u}(s)\leq vs-a_{t}+a_{t-s}+K,s\leq T(u)\right\}}\right)\leq C_{A,K}e^{-\delta R}.
Proof.

Let A,K,R>0A,K,R>0, we set

Yt(A,K,R)=u𝒩t2𝟙{Xu(t)mt(I)A}𝟙{T(u)tR}𝟙{Xu(s)vsat+ats+K,sT(u)}.Y_{t}(A,K,R)=\sum_{u\in\mathcal{N}^{2}_{t}}\mathbbm{1}_{\left\{X_{u}(t)\geq m^{(I)}_{t}-A\right\}}\mathbbm{1}_{\left\{T(u)\leq t-R\right\}}\mathbbm{1}_{\left\{X_{u}(s)\leq vs-a_{t}+a_{t-s}+K,s\leq T(u)\right\}}.

We use Proposition 4.1 to compute the mean of Yt(A,K,R)Y_{t}(A,K,R) as

𝐄(Yt(A,K,R))\displaystyle\mathbf{E}\left(Y_{t}(A,K,R)\right)\leq 0tReβs+ts𝐏(σBs+BtBsmt(I)A,σBrvrat+atr+K,rs)ds\displaystyle\int_{0}^{t-R}e^{\beta s+t-s}\mathbf{P}\left(\sigma B_{s}+B_{t}-B_{s}\geq m^{(I)}_{t}-A,\sigma B_{r}\leq vr-a_{t}+a_{t-r}+K,r\leq s\right)\mathrm{d}s
\displaystyle\leq 0tR𝐄(eθσBsF(ts,σBsat)𝟙{σBratrat+K,rs})ds,\displaystyle\int_{0}^{t-R}\mathbf{E}\left(e^{-\theta\sigma B_{s}}F(t-s,\sigma B_{s}-a_{t})\mathbbm{1}_{\left\{\sigma B_{r}\leq a_{t-r}-a_{t}+K,r\leq s\right\}}\right)\mathrm{d}s,

using the Markov property at time ss and the Girsanov transform, where F(r,x)=er𝐏(Brvrx)F(r,x)=e^{r}\mathbf{P}(B_{r}\geq vr-x). By the exponential Markov inequality, for all λ>0\lambda>0, we have F(r,x)eλxer(1λv+λ22)F(r,x)\leq e^{\lambda x}e^{r\left(1-\lambda v+\tfrac{\lambda^{2}}{2}\right)}. This implies

𝐄(Yt(A,K,R))0tRe(ts)(1λv+λ22)(t+1)3λ2θ𝐄(eσ(λθ)Bs𝟙{σBratrat+K,rs})ds.\mathbf{E}(Y_{t}(A,K,R))\leq\int_{0}^{t-R}e^{(t-s)(1-\lambda v+\tfrac{\lambda^{2}}{2})}(t+1)^{\tfrac{3\lambda}{2\theta}}\mathbf{E}\left(e^{\sigma(\lambda-\theta)B_{s}}\mathbbm{1}_{\left\{\sigma B_{r}\leq a_{t-r}-a_{t}+K,r\leq s\right\}}\right)\mathrm{d}s. (6.2)

We now bound this quantity in two different ways depending on the sign of σ21\sigma^{2}-1.

First, if σ2>1\sigma^{2}>1, then v>θv>\theta, in which case using (6.2) with λ=v\lambda=v, we obtain

𝐄(Yt(A,K,R))0tRe(ts)(1v2/2)(t+1)3v2θ𝐄(eσ(vθ)Bs𝟙{σBratrat+K,rs})ds.\mathbf{E}(Y_{t}(A,K,R))\leq\int_{0}^{t-R}e^{(t-s)(1-v^{2}/2)}(t+1)^{\tfrac{3v}{2\theta}}\mathbf{E}\left(e^{\sigma(v-\theta)B_{s}}\mathbbm{1}_{\left\{\sigma B_{r}\leq a_{t-r}-a_{t}+K,r\leq s\right\}}\right)\mathrm{d}s.

We now use that for all λ>0\lambda>0, there exists C>0C>0 such that for all 0st0\leq s\leq t, we have

𝐄(eλσBs𝟙{σBratrat+K,rs})CeλK(ts+1t+1)3λ2θ(s+1)32.\mathbf{E}\left(e^{\lambda\sigma B_{s}}\mathbbm{1}_{\left\{\sigma B_{r}\leq a_{t-r}-a_{t}+K,r\leq s\right\}}\right)\leq Ce^{\lambda K}\left(\tfrac{t-s+1}{t+1}\right)^{\frac{3\lambda}{2\theta}}(s+1)^{-\tfrac{3}{2}}. (6.3)

This bound can be obtained by classical Gaussian estimates, rewriting

𝐄(eλσBs𝟙{σBratrat+K,rs})CeλK(ts+1t+1)3λ2θk0eλk𝐏(σBsas+at+K[k1,k],σBratrat+K,rs),\mathbf{E}\left(e^{\lambda\sigma B_{s}}\mathbbm{1}_{\left\{\sigma B_{r}\leq a_{t-r}-a_{t}+K,r\leq s\right\}}\right)\\ \leq Ce^{\lambda K}\left(\tfrac{t-s+1}{t+1}\right)^{\frac{3\lambda}{2\theta}}\sum_{k\geq 0}e^{-\lambda k}\mathbf{P}(\sigma B_{s}-a_{s}+a_{t}+K\in[-k-1,-k],\sigma B_{r}\leq a_{t-r}-a_{t}+K,r\leq s),

and showing that the associated probability can be bounded uniformly in k,tk,t and sts\leq t by Ck(s+1)32Ck(s+1)^{-\tfrac{3}{2}}, with computations similar to the ones used in [Mal15a, Lemma 3.8] for random walks. Therefore, (6.3) implies that

𝐄(Yt(A,K,R))CA,K0tRe(ts)(1v2/2)(t+1)32(ts+1)3v2θ(s+1)32ds.\mathbf{E}(Y_{t}(A,K,R))\leq C_{A,K}\int_{0}^{t-R}e^{(t-s)(1-v^{2}/2)}\frac{(t+1)^{\tfrac{3}{2}}(t-s+1)^{\tfrac{3v}{2\theta}}}{(s+1)^{\tfrac{3}{2}}}\mathrm{d}s.

As (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, we have v=2βσ2>2v=\sqrt{2\beta\sigma^{2}}>\sqrt{2}, so 1v22<01-\tfrac{v^{2}}{2}<0. As a result

0t2e(ts)(1v2/2)(t+1)32(ts+1)3v2θ(s+1)32dsCe(1v2/2)t(t+1)5θ+3v2θ,\int_{0}^{\tfrac{t}{2}}e^{(t-s)(1-v^{2}/2)}\frac{(t+1)^{\tfrac{3}{2}}(t-s+1)^{\tfrac{3v}{2\theta}}}{(s+1)^{\tfrac{3}{2}}}\mathrm{d}s\leq Ce^{(1-v^{2}/2)t}(t+1)^{\tfrac{5\theta+3v}{2\theta}},

which converges to 0 as tt\to\infty, and

t2tRe(ts)(1v2/2)(t+1)32(ts+1)3v2θ(s+1)32dsCRe(1v2/2)s(s+1)3v2θdsCe(1v2/2)R/2.\int_{\tfrac{t}{2}}^{t-R}e^{(t-s)(1-v^{2}/2)}\frac{(t+1)^{\tfrac{3}{2}}(t-s+1)^{\tfrac{3v}{2\theta}}}{(s+1)^{\tfrac{3}{2}}}\mathrm{d}s\leq C\int_{R}^{\infty}e^{(1-v^{2}/2)s}(s+1)^{\tfrac{3v}{2\theta}}\mathrm{d}s\leq Ce^{(1-v^{2}/2)R/2}.

This completes the proof of the lemma in the case σ2>1\sigma^{2}>1.

We now assume that σ2<1\sigma^{2}<1. We have that

1θv+θ22=12β+βσ2=β(1σ22)+1.1-\theta v+\frac{\theta^{2}}{2}=1-2\beta+\frac{\beta}{\sigma^{2}}=\beta\left(\tfrac{1}{\sigma^{2}}-2\right)+1.

Therefore, as long as σ2>β2β1\sigma^{2}>\frac{\beta}{2\beta-1}, which is the case as σ2<1\sigma^{2}<1 and (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, we have 1θv+θ22<01-\theta v+\tfrac{\theta^{2}}{2}<0. Therefore, for all δ>0\delta>0 small enough such that

1(θ+δ)v+(θ+δ)22<0,1-(\theta+\delta)v+\tfrac{(\theta+\delta)^{2}}{2}<0,

using (6.2) with λ=θ+δ\lambda=\theta+\delta, we have

𝐄(Yt(A,K,R))0tRe(ts)(1(θ+δ)v+(θ+δ)2/2)(t+1)3(θ+δ)2θ𝐄(eσδBs𝟙{σBratrat+K,rs})ds.\mathbf{E}(Y_{t}(A,K,R))\leq\int_{0}^{t-R}e^{(t-s)(1-(\theta+\delta)v+(\theta+\delta)^{2}/2)}(t+1)^{\tfrac{3(\theta+\delta)}{2\theta}}\mathbf{E}\left(e^{\sigma\delta B_{s}}\mathbbm{1}_{\left\{\sigma B_{r}\leq a_{t-r}-a_{t}+K,r\leq s\right\}}\right)\mathrm{d}s.

So with the same computations as above, we obtain once again that

lim supt𝐄(Yt(A,K,R))CA,KeδR,\limsup_{t\to\infty}\mathbf{E}(Y_{t}(A,K,R))\leq C_{A,K}e^{-\delta R},

which completes the proof. ∎

Using the above computation, we deduce that with high probability, only particles of type 22 having an ancestor of type 11 at time tO(1)t-O(1) contribute substantially to the extremal process at time tt.

Lemma 6.2.

Assuming that (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, for all A>0A>0, we have

limRlim supt𝐏(u𝒩t2:T(u)tR,Xu(t)mt(I)A)=0.\lim_{R\to\infty}\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:T(u)\leq t-R,X_{u}(t)\geq m_{t}^{(I)}-A)=0.
Proof.

We observe that for all K>0K>0, we have

𝐏(u𝒩t2:T(u)tR,Xu(t)mt(I)A)𝐏(st,u𝒩s1:Xu(s)vs+atats)+𝐏(u𝒩t2:T(u)tR,Xu(t)mt(I)A,Xu(s)vsat+ats,sT(u)).\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:T(u)\leq t-R,X_{u}(t)\geq m_{t}^{(I)}-A)\leq\mathbf{P}(\exists s\leq t,u\in\mathcal{N}^{1}_{s}:X_{u}(s)\geq vs+a_{t}-a_{t-s})\\ +\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:T(u)\leq t-R,X_{u}(t)\geq m_{t}^{(I)}-A,X_{u}(s)\leq vs-a_{t}+a_{t-s},s\leq T(u)).

Then, using (6.1), the Markov inequality and Lemma 6.1, we obtain

lim supt𝐏(u𝒩t2:T(u)tR,Xu(t)mt(I)A)C(K+1)eθK+CA,KeR.\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:T(u)\leq t-R,X_{u}(t)\geq m_{t}^{(I)}-A)\leq C(K+1)e^{-\theta K}+C_{A,K}e^{-R}.

Letting RR\to\infty then KK\to\infty, the proof is now complete. ∎

For all R>0R>0, we set

^tR:=u𝒩t2𝟙{T(u)tR}δXu(t)mt(I).\widehat{\mathcal{E}}^{R}_{t}:=\sum_{u\in\mathcal{N}^{2}_{t}}\mathbbm{1}_{\left\{T(u)\geq t-R\right\}}\delta_{X_{u}(t)-m^{(I)}_{t}}.

We now show that ^tR\widehat{\mathcal{E}}^{R}_{t} converges in law as tt\to\infty.

Lemma 6.3.

Assume that (β,σ2)𝒞I(\beta,\sigma^{2})\in\mathcal{C}_{I}, there exists cR>0c_{R}>0 and a point measure distribution 𝒟R\mathcal{D}^{R} such that for all φ𝒯\varphi\in\mathcal{T}, we have

limt^tR,φ=^R,φ in law,\lim_{t\to\infty}{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}={\left\langle\widehat{\mathcal{E}}^{R}_{\infty},\varphi\right\rangle}\quad\text{ in law,}

where ^R\widehat{\mathcal{E}}^{R}_{\infty} is a DPPP(cRZeθxdxc_{R}Z_{\infty}e^{-\theta x}\mathrm{d}x,𝔇R\mathfrak{D}^{R}).

Proof.

We can rewrite

^tR=u𝒩tR1u𝒩t2uuδXu(t)Xu(tR)+Xu(tR)mt(I)=u𝒩tR1τXu(tR)mt(I)^R(u),\widehat{\mathcal{E}}^{R}_{t}=\sum_{u\in\mathcal{N}^{1}_{t-R}}\sum_{\begin{subarray}{c}u^{\prime}\in\mathcal{N}^{2}_{t}\\ u^{\prime}\succcurlyeq u\end{subarray}}\delta_{X_{u^{\prime}}(t)-X_{u}(t-R)+X_{u}(t-R)-m^{(I)}_{t}}=\sum_{u\in\mathcal{N}^{1}_{t-R}}\tau_{X_{u}(t-R)-m^{(I)}_{t}}\widehat{\mathcal{E}}^{(u)}_{R},

where τz\tau_{z} is the operator of translation by zz of point measures, and ^R(u)\widehat{\mathcal{E}}^{(u)}_{R} is the point process of descendants of type 22 of individual u𝒩tRu\in\mathcal{N}_{t-R} at time tt, centred around the position of uu at time tRt-R. Note that conditionally on tR1\mathcal{F}^{1}_{t-R}, (^R(u),u𝒩tR)(\widehat{\mathcal{E}}^{(u)}_{R},u\in\mathcal{N}_{t-R}) are i.i.d. point measures with same law as ¯R:=u𝒩R2δXu(R).\overline{\mathcal{E}}_{R}:=\sum_{u\in\mathcal{N}^{2}_{R}}\delta_{X_{u}(R)}.

Let φ𝒯\varphi\in\mathcal{T}, we set LL\in\mathbb{R} such that φ(x)=0\varphi(x)=0 for all xLx\leq L. By the branching property, we have

𝐄(e^tR,φ)=𝐄(u𝒩tR1FR(Xu(tR)mt(I)))=𝐄(eu𝒩tR1logFR(Xu(tR)mt(I))),\displaystyle\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(\prod_{u\in\mathcal{N}^{1}_{t-R}}F_{R}(X_{u}(t-R)-m^{(I)}_{t})\right)=\mathbf{E}\left(e^{-\sum_{u\in\mathcal{N}^{1}_{t-R}}-\log F_{R}(X_{u}(t-R)-m^{(I)}_{t})}\right),

where FR(x)=𝐄(exp(u𝒩R2φ(x+Xu(R))))F_{R}(x)=\mathbf{E}\left(\exp\left(-\sum_{u\in\mathcal{N}^{2}_{R}}\varphi(x+X_{u}(R))\right)\right). Observe that by Jensen transform, we have

logFR(x)\displaystyle-\log F_{R}(x) 𝐄(u𝒩R2φ(x+Xu(R)))0Reβs+(Rs)𝐄(φ(x+σBs+BtBs))ds\displaystyle\leq\mathbf{E}\left(\sum_{u\in\mathcal{N}^{2}_{R}}\varphi(x+X_{u}(R))\right)\leq\int_{0}^{R}e^{\beta s+(R-s)}\mathbf{E}(\varphi(x+\sigma B_{s}+B_{t}-B_{s}))\mathrm{d}s
φRe(β+1)R𝐏(B1xR(σ2+1)).\displaystyle\leq||\varphi||_{\infty}Re^{(\beta+1)R}\mathbf{P}\left(B_{1}\geq\tfrac{-x}{\sqrt{R(\sigma^{2}+1)}}\right).

Therefore, by (3.17), we have logFR(x)CRe(θ+δ)x1-\log F_{R}(x)\leq C_{R}e^{(\theta+\delta)x}\wedge 1 for all xx\in\mathbb{R}.

By Lemma 3.4, recall that u𝒩tR1δXu(tR)mt(I)\sum_{u\in\mathcal{N}^{1}_{t-R}}\delta_{X_{u}(t-R)-m^{(I)}_{t}} converges vaguely in law to a DPPP 1\mathcal{E}^{1} with intensity cθZeθ(z+vR)dzc_{\star}\theta Z_{\infty}e^{-\theta(z+vR)}\mathrm{d}z and decoration law 𝔇β,σ2\mathfrak{D}_{\beta,\sigma^{2}} the law of the decoration point measure of the BBM with branching rate β\beta and variance σ2\sigma^{2}. Additionally, it was proved by Madaule [Mad17] in the context of branching random walks, and extended in [CHL19] to BBM settings, that 1,eθ+δ<{\left\langle\mathcal{E}^{1},e_{\theta+\delta}\right\rangle}<\infty a.s. for all δ>0\delta>0, where eθ+δ(x)=e(θ+δ)xe_{\theta+\delta}(x)=e^{(\theta+\delta)x}. As a result, using the monotone convergence theorem, we obtain

limt𝐄(e^tR,φ)=𝐄(e1,logFR).\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(e^{-{\left\langle\mathcal{E}^{1},-\log F_{R}\right\rangle}}\right).

This proves that ^tR\widehat{\mathcal{E}}^{R}_{t} converges in law, as tt\to\infty, to a point process that can be obtained from 1\mathcal{E}^{1} by replacing each atom of 1\mathcal{E}^{1} by an independent copy of the point measure ¯R\overline{\mathcal{E}}_{R}. ∎

We now complete the proof of Theorem 1.1.

Proof of Theorem 1.1.

Let φ𝒯\varphi\in\mathcal{T}. We fix A>0A>0 such that φ(x)=0\varphi(x)=0 for all xAx\leq-A. We observe that

0𝐄(e^tR,φ)𝐄(e^t,φ)𝐏(u𝒩t2:T(u)tR,Xu(t)mt(I)A),0\leq\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)-\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)\leq\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:T(u)\leq t-R,X_{u}(t)\geq m_{t}^{(I)}-A),

which goes to 0 as tt then RR\to\infty, by Lemma 6.2. Additionally, by Lemma 6.3, we have

limt𝐄(e^tR,φ)=𝐄(e^R,φ).\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{\infty},\varphi\right\rangle}}\right).

Moreover, using that R𝐄(e^tR,φ)R\mapsto\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right) is decreasing, we deduce that

limt𝐄(e^t,φ)=limR𝐄(e^R,φ).\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)=\lim_{R\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{\infty},\varphi\right\rangle}}\right). (6.4)

Additionally, as R^tRR\mapsto\widehat{\mathcal{E}}^{R}_{t} is increasing in the space of point measures, we observe that we can construct the family of point measures (^R,R0)(\widehat{\mathcal{E}}^{R}_{\infty},R\geq 0) on the same probability space in such a way that almost surely, ^R,φ{\left\langle\widehat{\mathcal{E}}^{R}_{\infty},\varphi\right\rangle} is increasing for all φ\varphi. We denote by μ(φ)\mu(\varphi) its limit.

By [Kal02, Lemma 5.1], to prove that ^t\widehat{\mathcal{E}}_{t} admits a limit in distribution for the topology of vague convergence, it is enough to show that for all non-negative continuous functions with compact support, ^t,φ{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle} admits a limit in law which is a proper random variable. By (6.4), using the monotonicity of ^R\widehat{\mathcal{E}}^{R}_{\infty}, we immediately obtain that limt^t,φ=μ(φ)\lim_{t\to\infty}{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}=\mu(\varphi) in law. Therefore, to prove that ^t\widehat{\mathcal{E}}_{t} converges vaguely in distribution, it is enough to show that for all φ𝒯\varphi\in\mathcal{T}, μ(φ)<\mu(\varphi)<\infty a.s. which is a consequence of the tightness of ^t,φ{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}.

Let φ𝒯\varphi\in\mathcal{T}, we write LL\in\mathbb{R} such that φ(x)=0\varphi(x)=0 for all x<Lx<L. For all A>0A>0 and K>0K>0, we have

𝐏(^t,φA)𝐏(st,u𝒩s:Xu(s)vsat+ats+K)+1A𝐄(^t,φ𝟙{maxu𝒩s1Xu(s)vsat+ats+K,st}).\mathbf{P}\left({\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}\geq A\right)\leq\mathbf{P}(\exists s\leq t,\exists u\in\mathcal{N}_{s}:X_{u}(s)\geq vs-a_{t}+a_{t-s}+K)\\ +\frac{1}{A}\mathbf{E}\left({\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}\mathbbm{1}_{\left\{\max_{u\in\mathcal{N}^{1}_{s}}X_{u}(s)\leq vs-a_{t}+a_{t-s}+K,s\leq t\right\}}\right).

The first quantity goes to 0 as KK\to\infty by (6.1). Therefore, for all ε>0\varepsilon>0, we can fix KK large enough so that it remains smaller than ε/2\varepsilon/2. Then, using Lemma 6.1 with R=0R=0, we have

1A𝐄(^t,φ𝟙{maxu𝒩s1Xu(s)vsat+ats+K,st})CL,KA.\frac{1}{A}\mathbf{E}\left({\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}\mathbbm{1}_{\left\{\max_{u\in\mathcal{N}^{1}_{s}}X_{u}(s)\leq vs-a_{t}+a_{t-s}+K,s\leq t\right\}}\right)\leq\frac{C_{L,K}}{A}.

Therefore, we can choose AA large enough such that for all t0t\geq 0, 𝐏(^t,φA)ε\mathbf{P}({\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}\geq A)\leq\varepsilon, which completes the proof of the tightness of ^t,φ{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}.

We then conclude that ^t\widehat{\mathcal{E}}_{t} converges vaguely in law as tt\to\infty to a limiting point measure that we write ¯\overline{\mathcal{E}}. This point measure also is the limit as RR\to\infty of ^R\widehat{\mathcal{E}}^{R}_{\infty}, by (6.4). This allows us to show that ^t,φ¯,φ{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}\to{\left\langle\overline{\mathcal{E}},\varphi\right\rangle} in law for all φ𝒯\varphi\in\mathcal{T}, so we conclude by Proposition 3.3 that the position of the rightmost atom in ^t\widehat{\mathcal{E}}_{t} also converges to the position of the rightmost atom in ¯\overline{\mathcal{E}}.

To complete the proof of Theorem 1.1, we have to describe the law of ¯\overline{\mathcal{E}}. For all s0s\geq 0, using the branching property, we have

^t=u𝒩sτXu(s)vs+atsat^ts(u)\widehat{\mathcal{E}}_{t}=\sum_{u\in\mathcal{N}_{s}}\tau_{X_{u}(s)-vs+a_{t-s}-a_{t}}\widehat{\mathcal{E}}^{(u)}_{t-s}

where conditionally on s\mathcal{F}_{s}, (^ts(u),u𝒩s)(\widehat{\mathcal{E}}^{(u)}_{t-s},u\in\mathcal{N}_{s}) is a family of independent point measures with same law as ^ts\widehat{\mathcal{E}}_{t-s}, under law 𝐏(1)\mathbf{P}^{(1)} or 𝐏(2)\mathbf{P}^{(2)} depending on the type of uu. As no particle of type 22 born at an early time will have a descendant contributing in the extremal process by Lemma 6.2, we obtain that, letting tt\to\infty,

¯=(d)u𝒩s1τXu(s)vs¯(u),\overline{\mathcal{E}}{\overset{(d)}{=}}\sum_{u\in\mathcal{N}^{1}_{s}}\tau_{X_{u}(s)-vs}\overline{\mathcal{E}}^{(u)}, (6.5)

where ¯(u)\overline{\mathcal{E}}^{(u)} are i.i.d. copies of ¯\overline{\mathcal{E}}, that are further independent of s\mathcal{F}_{s}. This superposition property characterizes the law of ¯\overline{\mathcal{E}} as a decorated Poisson point process with intensity proportional to eθxdxe^{-\theta x}\mathrm{d}x, shifted by the logarithm of the derivative martingale of the branching Brownian motion by [Mai13, Corollary 3.2], with similar computations as in [Mad17, Section 2.2]. A general study of such point measures satisfying the branching property (6.5) is carried out in [MM21]. ∎

7 Asymptotic behaviour in the anomalous spreading case

We assume in this section that (σ2,β)𝒞III(\sigma^{2},\beta)\in\mathcal{C}_{III}, i.e. that β+σ2>2\beta+\sigma^{2}>2 and σ2<β2β1\sigma^{2}<\frac{\beta}{2\beta-1}. In particular, it implies that β>1\beta>1 and σ2<1\sigma^{2}<1. Under these conditions, we set

θ:=2β11σ2,a:=σ2θ,b:=θandp:=σ2+β22(β1)(1σ2),\theta:=\sqrt{2\frac{\beta-1}{1-\sigma^{2}}},\quad a:=\sigma^{2}\theta,\quad b:=\theta\quad\text{and}\quad p:=\frac{\sigma^{2}+\beta-2}{2(\beta-1)(1-\sigma^{2})},

which are the values of aa, bb and pp solutions of (2.1), described in terms of the parameter θ\theta which plays the role of a Lagrange multiplier in the optimization problem. Note that a<2βσ2a<\sqrt{2\beta\sigma^{2}}, b>2b>\sqrt{2} and p(0,1)p\in(0,1). Recall that in this situation, the maximal displacement is expected to satisfy

mt(III)=vt,where v=ap+b(1p)=βσ22(β1)(1σ2).m^{(III)}_{t}=vt,\quad\text{where }\ v=ap+b(1-p)=\frac{\beta-\sigma^{2}}{\sqrt{2(\beta-1)(1-\sigma^{2})}}.

As in the previous sections, we set

^t=u𝒩t2δXu(t)mt(III)\widehat{\mathcal{E}}_{t}=\sum_{u\in\mathcal{N}_{t}^{2}}\delta_{X_{u}(t)-m^{(III)}_{t}}

the appropriately centred extremal process of particles of type 22.

As mentioned in Section 2.1, under the above assumption, we are in the anomalous behaviour regime. In this regime, we have v>max(2,2βσ2)v>\max(\sqrt{2},\sqrt{2\beta\sigma^{2}}), in other words, this furthest particle travelled at a larger speed than the ones observed in the BBM of particles of type 11, or in a BBM of particles of type 22. Moreover, given the heuristic explanation for (2.1), we expect the furthest particle uu of type 22 at time tt to satisfy T(u)ptT(u)\approx pt and Xu(T(u))aptX_{u}(T(u))\approx apt.

The idea of the proof of Theorem 1.3 is to show that this heuristic holds, and that all particles participating to the extremal process of the multitype BBM are of type 22, and satisfy T(u)ptT(u)\approx pt and Xu(T(u))aptX_{u}(T(u))\approx apt. We then use the asymptotic behaviour of the growth rate of the number of particles of type 11 growing at speed aa to complete the proof. We begin by proving that with high probability, there is no particle of type 22 far above level mt(III)m^{(III)}_{t} at time tt.

Lemma 7.1.

Assuming that (σ2,β)𝒞III(\sigma^{2},\beta)\in\mathcal{C}_{III}, we have limAlim supt𝐏(u𝒩t2:Xu(t)mt(III)+A)=0.\displaystyle\lim_{A\to\infty}\limsup_{t\to\infty}\mathbf{P}\left(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}+A\right)=0.

Proof.

The proof of this result is based on a first moment method. For A>0A>0, we compute, using the many-to-one lemma, the mean of Xt(A)=u𝒩t2𝟙{Xu(t)mt(III)+A}.X_{t}(A)=\sum_{u\in\mathcal{N}_{t}^{2}}\mathbbm{1}_{\left\{X_{u}(t)\geq m_{t}^{(III)}+A\right\}}. Using (3.17), there exists C>0C>0 such that for all t1t\geq 1, we have

𝐄(Xt(A))\displaystyle\mathbf{E}(X_{t}(A)) =0teβs+ts𝐏(σBs+(BtBs)mt(III)+A)ds\displaystyle=\int_{0}^{t}e^{\beta s+t-s}\mathbf{P}\left(\sigma B_{s}+(B_{t}-B_{s})\geq m_{t}^{(III)}+A\right)\mathrm{d}s
0teβs+(ts)Cσ2s+ts(vt+A)e(vt+A)22(σ2s+ts)dsCt1/20teβs+(ts)e(vt+A)22(σ2s+ts)ds.\displaystyle\leq\int_{0}^{t}e^{\beta s+(t-s)}\frac{C\sqrt{\sigma^{2}s+t-s}}{(vt+A)}e^{-\frac{(vt+A)^{2}}{2(\sigma^{2}s+t-s)}}\mathrm{d}s\leq Ct^{-1/2}\int_{0}^{t}e^{\beta s+(t-s)}e^{-\frac{(vt+A)^{2}}{2(\sigma^{2}s+t-s)}}\mathrm{d}s.

Therefore, setting φ:uβu+1uv22(σ2u+1u)\varphi:u\mapsto\beta u+1-u-\frac{v^{2}}{2(\sigma^{2}u+1-u)}, by change of variable we have, for all tt large enough

𝐄(Xt(A))Ct1/201exp(tφ(u))eAv(σ2u+1u)du.\mathbf{E}(X_{t}(A))\leq Ct^{1/2}\int_{0}^{1}\exp\left(t\varphi(u)\right)e^{-A\frac{v}{(\sigma^{2}u+1-u)}}\mathrm{d}u.

We observe that

φ(u)=β1(1σ2)v22(σ2u+1u)2 and φ′′(u)=2(1σ2)2v22(σ2u+1u)3,\varphi^{\prime}(u)=\beta-1-(1-\sigma^{2})\frac{v^{2}}{2(\sigma^{2}u+1-u)^{2}}\quad\text{ and }\quad\varphi^{\prime\prime}(u)=-2(1-\sigma^{2})^{2}\frac{v^{2}}{2(\sigma^{2}u+1-u)^{3}},

hence φ\varphi is concave, and maximal at point u=pu=p, with a maximum equal to 0. By Taylor expansion, there exists δ>0\delta>0 such that φ(u)δ(up)2\varphi(u)\leq-\delta(u-p)^{2} for all u[0,1]u\in[0,1]. Therefore, we have

𝐄(Xt(A))CeAvt1/201eδ(up)2tduCeAvπ/δ.\mathbf{E}(X_{t}(A))\leq Ce^{-Av}t^{1/2}\int_{0}^{1}e^{-\delta(u-p)^{2}t}\mathrm{d}u\leq Ce^{-Av}\sqrt{\pi/\delta}.

As a result, applying the Markov inequality, we have

𝐏(u𝒩t2:Xu(t)mt(III)+A)=𝐏(Xt(A)1)𝐄(Xt(A)),\mathbf{P}\left(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}+A\right)=\mathbf{P}(X_{t}(A)\geq 1)\leq\mathbf{E}(X_{t}(A)),

thus there exists C>0C>0 such that

lim supt𝐏(u𝒩t2:Xu(t)mt(III)+A)CeAv,\limsup_{t\to\infty}\mathbf{P}\left(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}+A\right)\leq Ce^{-Av},

which converges to 0 as AA\to\infty. ∎

Next, we show that every particle of type 22 that contributes to the extremal process of the BBM branched from a particle of type 11 at a time and position close to (pt,apt)(pt,apt).

Lemma 7.2.

Assuming that (σ2,β)𝒞III(\sigma^{2},\beta)\in\mathcal{C}_{III}, for all A>0A>0, we have

limRlim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|T(u)pt|Rt1/2)=0,\displaystyle\lim_{R\to\infty}\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|T(u)-pt|\geq Rt^{1/2})=0, (7.1)
and limRlim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|Xu(T(u))apt|Rt1/2)=0.\displaystyle\lim_{R\to\infty}\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|X_{u}(T(u))-apt|\geq Rt^{1/2})=0. (7.2)
Proof.

Let A>0A>0 and ε>0\varepsilon>0. By Lemma 7.1, there exists K>0K>0 such that with probability (1ε)(1-\varepsilon) no particle of type 22 is above level mt(III)+Km^{(III)}_{t}+K at time tt for all tt large enough. For R>0R>0, we now compute the mean of

Yt(1)(A,K,R)=u𝒩t2𝟙{Xu(t)mt(III)[A,K]}𝟙{|T(u)pt|Rt1/2}.Y^{(1)}_{t}(A,K,R)=\sum_{u\in\mathcal{N}_{t}^{2}}\mathbbm{1}_{\left\{X_{u}(t)-m_{t}^{(III)}\in[-A,K]\right\}}\mathbbm{1}_{\left\{|T(u)-pt|\geq Rt^{1/2}\right\}}.

By Proposition 4.1, setting It(R)=[0,t]\[ptRt1/2,pt+Rt1/2]I_{t}(R)=[0,t]\backslash[pt-Rt^{1/2},pt+Rt^{1/2}] we have

𝐄(Yt(1)(A,K,R))\displaystyle\mathbf{E}\left(Y^{(1)}_{t}(A,K,R)\right) =It(R)eβs+(ts)𝐏(σBs+(BtBs)mt(III)[A,K])ds\displaystyle=\int_{I_{t}(R)}e^{\beta s+(t-s)}\mathbf{P}\left(\sigma B_{s}+(B_{t}-B_{s})-m_{t}^{(III)}\in[-A,K]\right)\mathrm{d}s
CeAvt1/2(0pRt1/2etφ(u)du+p+Rt1/21etφ(u)du),\displaystyle\leq Ce^{Av}t^{1/2}\left(\int_{0}^{p-Rt^{-1/2}}e^{t\varphi(u)}\mathrm{d}u+\int_{p+Rt^{-1/2}}^{1}e^{t\varphi(u)}\mathrm{d}u\right),

using the same notation and computation techniques as in the proof of Lemma 7.1. Thus, using again that there exists δ>0\delta>0 such that φ(u)δ(up)2\varphi(u)\leq-\delta(u-p)^{2} for some δ>0\delta>0, by change of variable z=t1/2(up)z=t^{1/2}(u-p) we obtain that

𝐄(Yt(1)(A,K,R))CeAv[R,R]eδz2dz.\mathbf{E}\left(Y^{(1)}_{t}(A,K,R)\right)\leq Ce^{Av}\int_{\mathbb{R}\setminus[-R,R]}e^{-\delta z^{2}}\mathrm{d}z.

Therefore, by Markov inequality, we obtain that

lim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|T(u)pt|Rt1/2)lim supt𝐏(u𝒩t2:Xu(t)mt(III)+K)+CeAv[R,R]eδz2dz.\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|T(u)-pt|\geq Rt^{1/2})\\ \leq\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}+K)+Ce^{Av}\int_{\mathbb{R}\setminus[-R,R]}e^{-\delta z^{2}}\mathrm{d}z.

As a result, with the choice previously made for the constant KK, we obtain that

lim supRlim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|T(u)pt|Rt1/2)ε.\limsup_{R\to\infty}\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|T(u)-pt|\geq Rt^{1/2})\leq\varepsilon.

By letting ε0\varepsilon\to 0, we complete the proof of (7.1).

We now turn to the proof of (7.2). By (7.1), we can assume, up to enlarging the value of KK that

lim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|T(u)pt|Kt1/2)ε.\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|T(u)-pt|\geq Kt^{1/2})\leq\varepsilon.

We now compute the mean of

Yt(2)(A,K,R)=u𝒩t2𝟙{Xu(t)mt(III)[A,K]}𝟙{|T(u)pt|Kt1/2}𝟙{|Xu(T(u))apt|Rt1/2}.Y^{(2)}_{t}(A,K,R)=\sum_{u\in\mathcal{N}_{t}^{2}}\mathbbm{1}_{\left\{X_{u}(t)-m_{t}^{(III)}\in[-A,K]\right\}}\mathbbm{1}_{\left\{|T(u)-pt|\leq Kt^{1/2}\right\}}\mathbbm{1}_{\left\{|X_{u}(T(u))-apt|\geq Rt^{1/2}\right\}}.

Using again Proposition 4.1, we have

𝐄(Yt(2)(A,K,R))\displaystyle\mathbf{E}\left(Y^{(2)}_{t}(A,K,R)\right)
=\displaystyle= ptKt1/2pt+Kt1/2eβs+ts𝐏(σBs+(BtBs)mt(III)[A,K],|σBsapt|Rt1/2)ds\displaystyle\int_{pt-Kt^{1/2}}^{pt+Kt^{1/2}}e^{\beta s+t-s}\mathbf{P}\left(\sigma B_{s}+(B_{t}-B_{s})-m_{t}^{(III)}\in[-A,K],|\sigma B_{s}-apt|\geq Rt^{1/2}\right)\mathrm{d}s
=\displaystyle= ptKt1/2pt+Kt1/2e2(β1)(spt)𝐄(eθ(σBs+BtBs)𝟙{σBs+(BtBs)+(ba)(pts)[A,K]|σBsa(pts)|Rt1/2})ds,\displaystyle\int_{pt-Kt^{1/2}}^{pt+Kt^{1/2}}e^{2(\beta-1)(s-pt)}\mathbf{E}\left(e^{\theta(\sigma B_{s}+B_{t}-B_{s})}\mathbbm{1}_{\left\{\begin{array}[]{l}\scriptstyle\sigma B_{s}+(B_{t}-B_{s})+(b-a)(pt-s)\in[-A,K]\\ \scriptstyle|\sigma B_{s}-a(pt-s)|\geq Rt^{1/2}\end{array}\right\}}\right)\mathrm{d}s,

by Girsanov transform, using that βs+tsθ22(σ2s+ts)=2(β1)(pts)\beta s+t-s-\frac{\theta^{2}}{2}(\sigma^{2}s+t-s)=2(\beta-1)(pt-s) and straightforward computations. Next, using that θ(ba)(pts)=2(β1)(spt)\theta(b-a)(pt-s)=-2(\beta-1)(s-pt), for RR large enough, we obtain

𝐄(Yt(2)(A,K,R))2Kt1/2eθKsup|r|Kt1/2𝐏(σBpt+r+BtBpt+r+(ba)r[A,K],|σBpt+rar|Rt1/2).\mathbf{E}\left(Y^{(2)}_{t}(A,K,R)\right)\\ \leq 2Kt^{1/2}e^{\theta K}\sup_{|r|\leq Kt^{1/2}}\mathbf{P}\left(\sigma B_{pt+r}+B_{t}-B_{pt+r}+(b-a)r\in[-A,K],|\sigma B_{pt+r}-ar|\geq Rt^{1/2}\right).

Then, by classical Gaussian computations, σBpt+r(σBpt+r+BtBpt+r)σ2pt+rσ2(pt+r)+t(1p)r\sigma B_{pt+r}-\left(\sigma B_{pt+r}+B_{t}-B_{pt+r}\right)\frac{\sigma^{2}pt+r}{\sigma^{2}(pt+r)+t(1-p)-r} is independent of σBpt+r+BtBpt+r\sigma B_{pt+r}+B_{t}-B_{pt+r}. We deduce that for RR large enough, we have for all tt large enough

sup|r|Kt1/2𝐏(σBpt+r+BtBpt+r+(ba)r[A,K],|σBpt+rar|Rt1/2)C(A+K)t1/2𝐏(|B1|R2(σ2p+1p)4σ2p(1p)).\sup_{|r|\leq Kt^{1/2}}\mathbf{P}\left(\sigma B_{pt+r}+B_{t}-B_{pt+r}+(b-a)r\in[-A,K],|\sigma B_{pt+r}-ar|\geq Rt^{1/2}\right)\\ \leq C(A+K)t^{-1/2}\mathbf{P}\left(|B_{1}|\geq\frac{R}{2}\sqrt{\tfrac{(\sigma^{2}p+1-p)}{4\sigma^{2}p(1-p)}}\right).

As a result, using again the Markov inequality, we have

lim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|Xu(T(u))apt|Rt1/2)lim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|T(u)pt|Kt1/2)+lim supt𝐏(u𝒩t2:Xu(t)mt(III)+K)+C(A+K)KeθK𝐏(|B1|R(σ2p+1p)4σ2p(1p)).\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|X_{u}(T(u))-apt|\geq Rt^{1/2})\\ \leq\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|T(u)-pt|\geq Kt^{1/2})\qquad\qquad\qquad\qquad\\ +\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}+K)+C(A+K)Ke^{\theta K}\mathbf{P}\left(|B_{1}|\geq R\sqrt{\tfrac{(\sigma^{2}p+1-p)}{4\sigma^{2}p(1-p)}}\right).

Hence, letting RR\to\infty, with the choice made for the constant KK, we obtain

lim supRlim supt𝐏(u𝒩t2:Xu(t)mt(III)A,|Xu(T(u))apt|Rt1/2)2ε,\limsup_{R\to\infty}\limsup_{t\to\infty}\mathbf{P}(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,|X_{u}(T(u))-apt|\geq Rt^{1/2})\leq 2\varepsilon,

and letting ε0\varepsilon\to 0 completes the proof of (7.2). ∎

The above lemma shows that typical particles of type 22 that contribute to the extremal process of the multitype BBM have their last ancestor of type 11 around time ptpt and position patpat. We now prove Theorem 1.3, using this localization of birth times and positions of particles in \mathcal{B} that have a descendant contribution to the extremal process at time tt, with high probability. Then, using Lemmas 3.1 and 4.2 we compute the quantity of contributing particles and with Lemma 3.7 to obtain the value associated to each contribution.

Proof of Theorem 1.3.

Let R>0R>0, we set

^tR:=u𝒩t2𝟙{|T(u)pt|Rt1/2,|Xu(T(u))apt|Rt1/2}δXu(t)mt(III).\widehat{\mathcal{E}}^{R}_{t}:=\sum_{u\in\mathcal{N}^{2}_{t}}\mathbbm{1}_{\left\{|T(u)-pt|\leq Rt^{1/2},|X_{u}(T(u))-apt|\leq Rt^{1/2}\right\}}\delta_{X_{u}(t)-m^{(III)}_{t}}.

Lemma 7.2 states that the extremal process ^tR\widehat{\mathcal{E}}^{R}_{t} is close to the extremal process of the BBM. Precisely, for all φ𝒯\varphi\in\mathcal{T} we have

|𝐄(e^tR,φ)𝐄(e^t,φ)|𝐏(u𝒩t2:Xu(t)mt(III)A,(T(u)pt,Xu(T(u))apt)[Rt1/2,Rt1/2]2)\left|\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)-\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)\right|\\ \leq\mathbf{P}\left(\exists u\in\mathcal{N}^{2}_{t}:X_{u}(t)\geq m_{t}^{(III)}-A,\ (T(u)-pt,X_{u}(T(u))-apt)\not\in[-Rt^{1/2},Rt^{1/2}]^{2}\right)

where AA is such that the support of φ\varphi is contained in [A,)[-A,\infty). As a result, by Lemma 7.2 we have

limRlim supt|𝐄(e^tR,φ)𝐄(e^t,φ)|=0,\lim_{R\to\infty}\limsup_{t\to\infty}\left|\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)-\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)\right|=0, (7.3)

so to compute the asymptotic behaviour of 𝐄(e^t,φ)\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right), it is enough to study the convergence of ^tR\widehat{\mathcal{E}}^{R}_{t} as tt then RR grow to \infty.

Let R>0R>0 and φ𝒯\varphi\in\mathcal{T}. Using the branching property and Corollary 4.3, we have

𝐄(e^tR,φ)=𝐄(exp(αptRt1/2pt+Rt1/2u𝒩s1𝟙{|Xu(s)apt|Rt1/2}F(ts,Xu(s)apt)ds)),\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(\exp\left(-\alpha\int_{pt-Rt^{1/2}}^{pt+Rt^{1/2}}\sum_{u\in\mathcal{N}^{1}_{s}}\mathbbm{1}_{\left\{|X_{u}(s)-apt|\leq Rt^{1/2}\right\}}F(t-s,X_{u}(s)-apt)\mathrm{d}s\right)\right),

with F(r,x)=1𝐄(2)(eu𝒩r2φ(Xu(r)br+xb((1p)tr)))F(r,x)=1-\mathbf{E}^{(2)}\left(e^{-\sum_{u\in\mathcal{N}^{2}_{r}}\varphi(X_{u}(r)-br+x-b((1-p)t-r))}\right). Additionally, by Lemma 3.7, we have

F((1p)tr,x)\displaystyle F((1-p)t-r,x) =C(b)e(1b22)((1p)tr)2πt(1p)eb(xbr)(xbr)22((1p)tr)eθz(1eΨb[φ](z))dz(1+o(1))\displaystyle=C(b)\frac{e^{(1-\frac{b^{2}}{2})((1-p)t-r)}}{\sqrt{2\pi t(1-p)}}e^{b(x-br)-\tfrac{(x-br)^{2}}{2((1-p)t-r)}}\int e^{-\theta z}\left(1-e^{-\Psi^{b}[\varphi](z)}\right)\mathrm{d}z(1+o(1))
=C(b)e(1b22)(1p)t2πt(1p)eθx(1+θ22)r(xbr)22(1p)teθz(1eΨb[φ](z))dz(1+o(1)),\displaystyle=C(b)\frac{e^{(1-\frac{b^{2}}{2})(1-p)t}}{\sqrt{2\pi t(1-p)}}e^{\theta x-(1+\frac{\theta^{2}}{2})r-\tfrac{(x-br)^{2}}{2(1-p)t}}\int e^{-\theta z}\left(1-e^{-\Psi^{b}[\varphi](z)}\right)\mathrm{d}z(1+o(1)),

as tt\to\infty, uniformly in |r|Rt1/2|r|\leq Rt^{1/2} and |x|Rt1/2|x|\leq Rt^{1/2}, where we used that θ=b\theta=b. Thus, setting

Θ(φ):=αC(b)eθz(1eΨb[φ](z))dz\Theta(\varphi):=\alpha C(b)\int e^{-\theta z}\left(1-e^{-\Psi^{b}[\varphi](z)}\right)\mathrm{d}z

and GR(r,x)=𝟙{|x+ar|R}e(x+(ab)r)22(1p)G_{R}(r,x)=\mathbbm{1}_{\left\{|x+ar|\leq R\right\}}e^{-\frac{(x+(a-b)r)^{2}}{2(1-p)}} we can now rewrite the Laplace transform of ^tR\widehat{\mathcal{E}}^{R}_{t} as

𝐄(e^tR,φ)\displaystyle\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)
=\displaystyle= 𝐄(exp(Θ(φ)(1+o(1))e(1b22)(1p)t2πt(1p)ptRt1/2pt+Rt1/2u𝒩s1eθ(Xu(s)apt)(1+θ22)(spt)GR(sptt1/2,Xu(s)ast1/2)ds)).\displaystyle\mathbf{E}\left(\exp\left(-\Theta(\varphi)(1+o(1))\tfrac{e^{(1-\frac{b^{2}}{2})(1-p)t}}{\sqrt{2\pi t(1-p)}}\!\!\int_{pt-Rt^{1/2}}^{pt+Rt^{1/2}}\!\!\!\!\sum_{u\in\mathcal{N}^{1}_{s}}e^{\theta(X_{u}(s)-apt)-(1+\frac{\theta^{2}}{2})(s-pt)}G_{R}\left(\tfrac{s-pt}{t^{1/2}},\tfrac{X_{u}(s)-as}{t^{1/2}}\right)\mathrm{d}s\right)\right).

By definition of the parameters, we have β+θ2σ22=1+θ22=βσ21σ2\beta+\frac{\theta^{2}\sigma^{2}}{2}=1+\frac{\theta^{2}}{2}=\frac{\beta-\sigma^{2}}{1-\sigma^{2}} and θa=β+θ2σ22(βa22σ2)\theta a=\beta+\frac{\theta^{2}\sigma^{2}}{2}-\left(\beta-\frac{a^{2}}{2\sigma^{2}}\right), therefore we can rewrite

𝐄(e^tR,φ)=𝐄(exp(Θ(φ)1+o(1)2πt(1p)ptRt1/2pt+Rt1/2u𝒩s1eθXu(s)(β+θ2σ22)sGR(sptt1/2,Xu(s)ast1/2)ds)),\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)\\ =\mathbf{E}\left(\exp\left(-\Theta(\varphi)\frac{1+o(1)}{\sqrt{2\pi t(1-p)}}\int_{pt-Rt^{1/2}}^{pt+Rt^{1/2}}\sum_{u\in\mathcal{N}^{1}_{s}}e^{\theta X_{u}(s)-(\beta+\frac{\theta^{2}\sigma^{2}}{2})s}G_{R}\left(\tfrac{s-pt}{t^{1/2}},\tfrac{X_{u}(s)-as}{t^{1/2}}\right)\mathrm{d}s\right)\right), (7.4)

where we used that (1p)(1b22)+p(1a22σ2)=0(1-p)(1-\frac{b^{2}}{2})+p(1-\frac{a^{2}}{2\sigma^{2}})=0.

We now observe that by Lemma 3.1, using (3.1), we have

limt1t1/2ptRt1/2pt+Rt1/2u𝒩s1eθXu(s)(β+θ2σ22)sGR(sptt1/2,Xu(s)aptt1/2)ds=W(θ)2πpσ2[R,R]×ez22σ2pe(z+(ab)r)22(1p)𝟙{|z+ar|R}drdz\lim_{t\to\infty}\frac{1}{t^{1/2}}\int_{pt-Rt^{1/2}}^{pt+Rt^{1/2}}\sum_{u\in\mathcal{N}^{1}_{s}}e^{\theta X_{u}(s)-(\beta+\frac{\theta^{2}\sigma^{2}}{2})s}G_{R}\left(\tfrac{s-pt}{t^{1/2}},\tfrac{X_{u}(s)-apt}{t^{1/2}}\right)\mathrm{d}s\\ =\frac{W_{\infty}(\theta)}{\sqrt{2\pi p\sigma^{2}}}\int_{[-R,R]\times\mathbb{R}}e^{-\frac{z^{2}}{2\sigma^{2}p}}e^{-\frac{(z+(a-b)r)^{2}}{2(1-p)}}\mathbbm{1}_{\left\{|z+ar|\leq R\right\}}\mathrm{d}r\mathrm{d}z

where W(θ)=limtu𝒩t1eθXu(t)t(β+σ2θ22)W_{\infty}(\theta)=\lim_{t\to\infty}\sum_{u\in\mathcal{N}^{1}_{t}}e^{\theta X_{u}(t)-t(\beta+\frac{\sigma^{2}\theta^{2}}{2})} is the limit of the additive martingale with parameter θ\theta for the branching Brownian motion of type 11. As a result, writing

cR=12πp(1p)σ2[R,R]×ez22σ2pe(z+(ab)r)22(1p)𝟙{|z+ar|R}drdz(0,),c_{R}=\frac{1}{2\pi\sqrt{p(1-p)\sigma^{2}}}\int_{[-R,R]\times\mathbb{R}}e^{-\frac{z^{2}}{2\sigma^{2}p}}e^{-\frac{(z+(a-b)r)^{2}}{2(1-p)}}\mathbbm{1}_{\left\{|z+ar|\leq R\right\}}\mathrm{d}r\mathrm{d}z\in(0,\infty),

by dominated convergence theorem, (7.4) yields

limt𝐄(e^tR,φ)=𝐄(exp(cRW(θ)Θ(φ))).\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}^{R}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(\exp\left(-c_{R}W_{\infty}(\theta)\Theta(\varphi)\right)\right).

This convergence holds for all φ𝒯\varphi\in\mathcal{T}. Then by [BBCM20, Lemma 4.4], the process ^tR\widehat{\mathcal{E}}_{t}^{R} converges vaguely in distribution as tt\to\infty to a DPPP(θcRW(θ)eθzdz,𝔇b\theta{c}_{R}W_{\infty}(\theta)e^{-\theta z}\mathrm{d}z,\mathfrak{D}^{b}), as tt\to\infty, where 𝔇b\mathfrak{D}^{b} is the law of 𝒟b\mathcal{D}^{b}, the point measure defined in (3.19).

To complete the proof, we now observe that by monotone convergence theorem, we have

limRcR=12πp(1p)σ2×ez22σ2pe(z+(ab)r)22(1p)dr=1ba=1θ(1σ2).\lim_{R\to\infty}c_{R}=\frac{1}{2\pi\sqrt{p(1-p)\sigma^{2}}}\int_{\mathbb{R}\times\mathbb{R}}e^{-\frac{z^{2}}{2\sigma^{2}p}}e^{-\frac{(z+(a-b)r)^{2}}{2(1-p)}}\mathrm{d}r=\frac{1}{b-a}=\frac{1}{\theta(1-\sigma^{2})}.

Therefore, letting tt\to\infty then RR\to\infty, (7.3) yields

limt𝐄(e^t,φ)=𝐄(exp(αC(b)W(θ)2(β1)θeθz(1eΨb[φ](z)))).\lim_{t\to\infty}\mathbf{E}\left(e^{-{\left\langle\widehat{\mathcal{E}}_{t},\varphi\right\rangle}}\right)=\mathbf{E}\left(\exp\left(-\frac{\alpha C(b)W_{\infty}(\theta)}{2(\beta-1)}\int\theta e^{-\theta z}\left(1-e^{-\Psi^{b}[\varphi](z)}\right)\right)\right).

As a result, using [BBCM20, Lemma 4.4], the proof of Theorem 1.3 is now complete, with c(III)=αC(b)2(β1)c^{(III)}=\frac{\alpha C(b)}{2(\beta-1)} and 𝔇(III)\mathfrak{D}^{(III)} the law of 𝒟b\mathcal{D}^{b} defined in (3.19). ∎

Acknowledgements.

The authors are partially funded by ANR-16-CE93-0003 (ANR MALIN). Additionally, M.A.B. is partially supported by Cofund MathInParis project from FSMP.

[Uncaptioned image]

This program has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 754362.

References