Let {Z(t);t≥0} be a continuous-time supercritical branching process with immigration (MBPI) with the offspring mean m(t). In this paper, we mainly research the lower deviation probabilities P(Z(t)=kt) and P(0≤Z(t)≤kt) with kt/em(t)→0 as t→∞. Moreover, we present the local limit theorem and some related estimates of the MBPIs. For our proofs, we use the well-known Cramér method to prove the large deviation of the sum of independent variables to satisfy our needs.
Citation: Juan Wang, Chao Peng. Lower deviation probabilities for supercritical Markov branching processes with immigration[J]. AIMS Mathematics, 2025, 10(5): 10324-10339. doi: 10.3934/math.2025470
[1] | You Lv . Asymptotic behavior of survival probability for a branching random walk with a barrier. AIMS Mathematics, 2023, 8(2): 5049-5059. doi: 10.3934/math.2023253 |
[2] | Xiangqi Zheng . On the extinction of continuous-state branching processes in random environments. AIMS Mathematics, 2021, 6(1): 156-167. doi: 10.3934/math.2021011 |
[3] | István Fazekas, Attila Barta, László Fórián, Bettina Porvázsnyik . A continuous-time network evolution model describing $ N $-interactions. AIMS Mathematics, 2024, 9(12): 35721-35742. doi: 10.3934/math.20241695 |
[4] | Guowei Zhang . The lower bound on the measure of sets consisting of Julia limiting directions of solutions to some complex equations associated with Petrenko's deviation. AIMS Mathematics, 2023, 8(9): 20169-20186. doi: 10.3934/math.20231028 |
[5] | Chahn Yong Jung, Muhammad Shoaib Saleem, Shamas Bilal, Waqas Nazeer, Mamoona Ghafoor . Some properties of η-convex stochastic processes. AIMS Mathematics, 2021, 6(1): 726-736. doi: 10.3934/math.2021044 |
[6] | Jawdat Alebraheem, Mogtaba Mohammed, Ismail M. Tayel, Muhamad Hifzhudin Noor Aziz . Stochastic prey-predator model with small random immigration. AIMS Mathematics, 2024, 9(6): 14982-14996. doi: 10.3934/math.2024725 |
[7] | Jawdat Alebraheem . Asymptotic stability of deterministic and stochastic prey-predator models with prey herd immigration. AIMS Mathematics, 2025, 10(3): 4620-4640. doi: 10.3934/math.2025214 |
[8] | Zongmin Yue, Yitong Li, Fauzi Mohamed Yusof . Dynamic analysis and optimal control of Zika virus transmission with immigration. AIMS Mathematics, 2023, 8(9): 21893-21913. doi: 10.3934/math.20231116 |
[9] | Hongwu Li, Yuling Feng, Hongwei Jiao, Youlin Shang . A novel algorithm for solving sum of several affine fractional functions. AIMS Mathematics, 2023, 8(4): 9247-9264. doi: 10.3934/math.2023464 |
[10] | Qingwu Gao, Wenlei Pan . Precise large deviations of aggregate claims in a nonstandard risk model with arbitrary dependence between claim sizes and waiting times. AIMS Mathematics, 2023, 8(1): 2191-2200. doi: 10.3934/math.2023113 |
Let {Z(t);t≥0} be a continuous-time supercritical branching process with immigration (MBPI) with the offspring mean m(t). In this paper, we mainly research the lower deviation probabilities P(Z(t)=kt) and P(0≤Z(t)≤kt) with kt/em(t)→0 as t→∞. Moreover, we present the local limit theorem and some related estimates of the MBPIs. For our proofs, we use the well-known Cramér method to prove the large deviation of the sum of independent variables to satisfy our needs.
Consider a continuous-time Markov branching process with immigration (MBPI), denoted {Z(t);t≥0}. This process comprises two components, including existing individuals and external immigration. Moreover, these two components are independently and identically distributed, following the evolutionary law determined by the branching rates {bj;j≥0,j≠1} and the immigration rates {aj;j≥1}
{bj≥0(j≠1),0<−b1=∑j≠1bj<∞,aj≥0(j≠0),0<−a0=∑j≠0aj<∞, | (1.1) |
respectively. The corresponding Q-matrix Q={qij;i,j∈Z+} is defined as follows:
qij:={ibj−i+1+aj−i if i≥0,j≥i,ib0 if i≥0,j=i−1,0 otherwise. | (1.2) |
Hence, the process above is completely determined by the infinitesimal generating functions B(u)=∑∞j=0bjuj and A(u)=∑∞j=0ajuj for u∈[0,1).
Throughout this paper, we assume that b0=0 and m:=∑∞j=0jbj<∞. By the definition of the branching rates {bj;j≥0} and the fact that b0=0, we can deduce that m>0, indicating that the Markov branching process is supercritical.
If a0=0, then Z(t) represents a pure branching process defined by {Z0(t);t≥0}. From Athreya and Ney[1], we know that a normalized function C(t) exists such that W(t):=Z0(t)/C(t)→W as t tends to ∞, where W is a nondegenerate random variable and C(t) satisfies limt→∞C(t+s)/C(t)=E[Z0(s)]=ems. Moreover, we find that EW=1 holds if and only if the LlogL-moment condition holds. If this condition holds, W has a continuous density function w(y) on (0,∞) such that the following global limit theorem holds:
limt→∞P(W(t)≥x)=∫∞xw(y)dy, x>0. | (1.3) |
If a0≠0, according to Li et al. [2], we see that Z(t+s)/Z(t) converges to ems in probability. Note that Z(t) can be expressed as
Z(t+s)=Z(t)∑i=1ξt,i(s)+Y(t):=Z0(t)+Y(t), | (1.4) |
where {ξt,i(s);t≥0,i≥1} are independently and identically distributed (i.i.d.) random variables with the same law as Z0(s) and Y(t) is the number of particles at moment t+s, being either immigrants or offspring of immigrants in (t,t+s], which is independent of {ξt,i(s);i≥1} and Z(t). Obviously, the distribution of Y(⋅) is independent of t.
Here, we are interested in the lower deviation probabilities P(Z(t)=kt) and P(0≤Z(t)≤kt) as kt/C(t)→0 (t→∞), since these probabilities characterize the evolution of the population when the population growth is below the average growth rate. Besides being of some interest in its own right, the asymptotic behavior of these probabilities is related to large deviations.
According to previous literature, large deviations are important and has drawn widespread attention from scholars. For the supercritical branching process, Athreya and Ney [3] considered the local limit theorem and some related aspects. Athreya [4] discussed the decay rates of P(|Zn+1/Zn−λ|>ε)(where λ denotes the offspring's mean) for a classical Galton Watson process {Zn;n≥1}. Ney and Vidyashankar [5] considered the harmonic moments and large deviation rates in 2003. Moreover, Ney and Vidyashankar [6] considered the local limit theory and large deviations in 2004. Fleischmann and Wachtel [7] studied the lower deviation probabilities.
For the supercritical branching processes with immigration {Xn;n≥1}, Seneta[8] and Pakes[9] considered the supercritical Galton Watson process with immigration. Chu et al. [10] researched the small value probabilities. Liu and Zhang [11] studied the decay rates of P(|Xn+1/Xn−λ|>ε). Sun and Zhang [12,13] considered the convergence rates of harmonic moments and the lower deviations. Furthermore, Li and Zhang [14] focused on the harmonic moments and large deviations for a critical Galton Watson process with immigration.
In recent years, the continuous-time Markov branching processes have drawn widespread attention. For example, Li et al.[15] researched the large deviation rates for Markov branching processes. For the Markov branching process with immigration, based on Li et al.[15], Li et al.[16] studied asymptotic properties. Inspired by [7] and [2], we deal with the asymptotic behavior of P(Z(t)=kt) and P(0≤Z(t)≤kt), known as global and local lower deviation probabilities under the assumption of E[Z(1)logZ(1)]<∞. Moreover, we apply the Cramér method to analyze the large deviation of the sum of independent variables.
The rest of this paper is organized as follows. In Section 2, we state some necessary preliminaries, apply the Cramér method to the MBPIs, and present some related estimates. In Section 3, we list the main results. Section 4 is devoted to concrete and detailed proofs concerning the main results.
Define P0(t)=(p0ij(t);i≥1,j≥1) as the transition function of the pure branching process without immigration {Z0(t);t≥0} and let F0(s,t)=∑∞j=0p01j(t)sj be the probability generating function of {Z0(t);t≥0} with the initial state Z0(0)=1.
Let P(t)=(pij(t);i≥1,j≥1) be the transition function of the MBPI {Z(t);t≥0} and Gl(s,t):=E[sZ(t)|Z(0)=l]=∑∞j=0plj(t)sj with Gl(s,0)=sl for 0≤s<1. Moreover, by Li et al. [16]
Gl(s,t)=H(s,t)⋅[F0(s,t)]l, l∈Z+, | (2.1) |
where H(s,t)=G0(s,t). In particular, G(s,t):=G1(s,t)=H(s,t)F0(s,t) with the initial state Z(0)=1.
Proposition 2.1. Define the function Q(v)=∑∞j=1qjvj as the unique solution of
B(v)Q′(v)+(A(v)−a0−b1)Q(v)=0,0⩽v<1, |
subject to
Q(0)=0, Q′(0)=1, Q(1)=∞ and Q(v)<∞ for 0⩽v<1, |
where qj satisfies
qj:={p11(t)e−(b1+a0)t if j=1,limt→∞p1j(t)e−(b1+a0)t if j≥2, |
with q1=1, qj⩽Πj−1k=1(1+a0kb1) (j≥2).
Proof. This follows from the Kolmogorov forward equation.
Proposition 2.2. For any 0≤s<1 and t>0, take
R(s,t):=H(s,t)ea0t, Q0l(s,t):=F0(s,t)eb1lt. | (2.2) |
Then
Ql(s,t):=Gl(s,t)e(a0+b1l)t=R(s,t)(Q0(s,t))l↗R(s)(Q0(s))l=:Ql(s), t→∞, | (2.3) |
where R(s):=limt→∞R(s,t) and Q0(s):=limt→∞Q0(s,t).
Proof. By Liu and Zhang[11], together with the definition of G(⋅,⋅), we can derive the results above.
For convenience, we always assume that the following conditions (A1)−(A3) hold throughout this paper:
(A1) b0=0;
(A2) 0<m<∞;
(A3) E[Z(1)logZ(1)]<+∞.
On this basis, we give the main results of this paper in Section 3. First, we will introduce the Cramér transformation, which is the most critical step in the famous Cramér method (see Petrov [17]). By this transformation, we can obtain some important related estimates for the subsequent proofs.
For the real random variable X, let X(h) be the random variable resulting from the Cramér transform determined by the constant h∈R. Then X(h) satisfies the following:
E[eiaX(h)]=E[e(h+ia)X]E[ehX], a∈R, | (2.4) |
where E[ehX]<∞.
Set the random variable X=Z(t). Following the equation (2.4) above, it can be inferred that for any h≤0, the Cramér transformation Z(−h/emt) exists and satisfies
E[eiaZ(−h/emt)]=Gl(e−h/emt+ia,t)Gl(e−h/emt,t), a∈R. | (2.5) |
For the convenience of discussion, we first give the Cramér transformation of Z0(t) and Y(t) separately, since the branching part Z0(t) and the pure immigration part Y(t) are independent. For any h≤0 and t≥0, we define a sequence of random variables {Xi(h,t);i≥1}, which are i.i.d. according to the Cramér transform of Z0(t) determined by the constant −h/emt, i.e.,
P(X1(h,t)=k)=e−kh/emtF0(e−h/emt,t)P(Z0(t)=k), k≥1. |
The equation above can be rewritten as
E[eiaZ0(−h/emt)]=E[eiaX1(h,t)]=F0(e−h/emt+ia,t)F0(e−h/emt,t). |
For the pure immigrant part Y(t), the Cramér transform is determined by the constant −h/emt. We can define a random variable T(h,t) that is independent of Xi(h,t) and Y(t). Thus, we have
P(T(h,t)=k)=e−kh/emtH(e−h/emt,t)P(Y(t)=k), k≥1, |
which is equivalent to
E[eiaT(h,t)]=E[eiaY(−h/emt)]=H(e−h/emt+ia,t)H(e−h/emt,t). |
After the transformations above, we construct a sequence of independent random variables {Sl(h,t);t≥0,l≥1} expressed as
Sl(h,t):=l∑i=1Xi(h,t)+T(h,t), l≥1. | (2.6) |
Thus
P(Sl(h,t)=k)=e−kh/emtGl(e−h/emt,t)P(Z(t)=k|Z(0)=l), k≥1. | (2.7) |
Proof. In order to prove that Sl(h,t) and Z(−h/emt,t) are identically distributed. It follows from (2.4) and the definition of Sl(h,t) that
EeiaSl(h,t)=Eeia[∑li=1Xi(h,t)+T(h,t)]=(EeiaZ0(−h/emt))l⋅EeiaY(−h/emt)=(Ee(−h/emt+ia)Z0(t)Ee(−h/emt)Z0(t))l⋅Ee(−h/emt+ia)Y(t)Ee(−h/emt)Y(t)=(F0(e−h/emt+ia,t)F0(e−h/emt,t))l⋅H(e−h/emt+ia,t)H(e−h/emt,t)=Gl(e−h/emt+ia,t)Gl(e−h/emt,t). |
On the other hand,
EeiaSl(h,t)=∞∑k=0eiakP(Sl(h,t)=k). |
If we compare the corresponding coefficients of the equations above, then (2.7) is established.
Denoting the characteristic function of Sl(h,t) as ΨS(a):=E[eiaSl(h,t)], then we obtain
ΨS(a)=Gl(e−h/emt+ia,t)Gl(e−h/emt,t). | (2.8) |
By comparing (2.8) and (2.5), it can be found that Sl(h,t) and Z(−h/emt,t) are identically distributed.
Definition 2.1. (Concentration function, [17], p. 38) For any λ≥0, the concentration function Q(X;λ) of a random variable X is defined by the equality
Q(X;λ):=supxP(x≤X≤x+λ). | (2.9) |
Lemma 2.1. (Petrov[17], p. 38, Lemma 3) Suppose that X is a random variable with the characteristic function ψ(t). Then for λ≥0 and θ>0, we have
Q(X;λ)≤(9695)2max(λ,1θ)∫θ−θ|ψ(t)|dt. | (2.10) |
Lemma 2.2. For any h≥0, C(h) exists such that
supt,k≥0emtP(Sl(h,t)=k)≤C(h)l−1/2, l≥l0:=1+[1/α], | (2.11) |
where α=−logσm and σ:=∂G(u,t)∂u|(0,1)=p11(1).
Proof. First, if we recall Proposition 2.1, it is not difficult to find p11(t)=e(b1+a0)t. By Lemma 2.2, we also have σ:=∂G(u,t)∂u|(0,1)=p11(1), and then we get σ=eb1+a0 and thus it satisfies p11(t)=σt.
First, we prove that (2.11) holds when l=l0=[1/α]+1. Taking X=Sl0(h,t), λ=1/2, and θ=π in (2.10) together with (2.8), we have
Q(Sl0(h,t);12)≤C∫π−π|Gl0(e−h/emt+ia,t)|Gl0(e−h/emt,t)da, | (2.12) |
where C is a constant independent of h and t. Note that Sl0(h,t) is a non-negative integer random variable, then from (2.9), the inequality above is equivalent to
supk≥1P(Sl0(h,t)=k)≤C∫π−π|Gl0(e−h/emt+ia,t)|Gl0(e−h/emt,t)da. | (2.13) |
Consider the denominator part of the inequality above, owing to (1.4) and the branching property, we have
Gl0(e−h/emt,t)=(E[e−h(Z0(t)/emt)])l0(E[e−h(Y(t)/emt)])t→∞⟶E[e−h(W∗l0+I)]>0, | (2.14) |
where W∗l0 represents the l-fold convolution of W.
The convergence above is uniform for h in a compact subset of R+. Thus for any fixed h, a positive number t0 exists such that Gl0(e−h/emt,t)>δ>0 for all t≥t0, which implies inft>0Gl0(e−h/emt,t)>0. According to (2.12), a constant C(h) exists such that
supk≥1P(Sl0(h,t)=k)≤C(h)∫π−π|Gl0(e−h/emt+ia,t)|da. | (2.15) |
Now we only need to prove that the right side of the inequality above is bounded.
∫π−π|Gl0(e−h/emt+ia,t)|da=(∫−πe−mt−π+∫πe−mt−πe−mt+∫ππe−mt)|Gl0(e−h/emt+ia,t)|da:=I1+I2+I3. | (2.16) |
Clearly,
I2=∫πe−mt−πe−mt|Gl0(e−h/emt+ia,t)|da≤2πe−mt. | (2.17) |
It follows, by (2.3), that Gl(s,t)e(a0+b1l)t↑Ql(s) as t↑∞, and then
I3≤∫ππe−mtQl0(e−h/emt+ia)e(a0+b1l0)tda. | (2.18) |
It is easy to find that when t is large enough, a constant C1(h) exists such that the integrand is bounded. We can get I3≤πe−mte(a0+b1l0)tC1(h). Taking u=−a for I1, after a similar analysis, I1≤πe−mte(a0+b1l0)tC2(h) is obatined. Then C3(h) exists such that C1(h)+C2(h)≤2C3(h). Hence
I1+I2+I3≤2πe−mt[e(a0+b1l0)tC3(h)+1]. | (2.19) |
By the definition of a0, b1, and m, the right side of (2.19) is bounded in t∈(0,∞). Due to (2.13), one has
supt≥0,k≥1emtP(Sl(h,t)=k)≤C(h). | (2.20) |
When l>l0, the result also holds as shown below. The sequence {Xi(h,t)}i≥1 given by the Cramér transformation is a sequence of independently and identically distributed random variables. Set S0l0(h,t):=∑l0i=1Xi(h,t), then S0l0(h,t) is also nondegenerate, since Xi(h,t) is nondegenerate. According to (98) in [7], for j≥1, D1(h)>0, and D2(h)>0, we have
supt≥0,k≥1emtP(S0jl0(h,t)=k)≤D1(h)√j=D2(h)√jl0. | (2.21) |
According to Lemma 1 from Chapter Ⅲ of Petrov [17], for any two independent random variables X and Y, the inequality Q(X+Y;λ)≤Q(min(X,Y);λ) holds for ∀ λ>0. Now, letting X=S0l(h,t) and Y=T(h,t), by (2.6), we have
supt≥0,k≥1emtP(Sl(h,t)=k)≤supt≥0,k≥1emtP(S0l(h,t)=k)≤supt≥0,k≥1emtP(S0[l/l0]l0(h,t)=k). | (2.22) |
Taking j=[l/l0]≥1 in (2.19), the inequality (2.11) above also holds for every l>l0.
Lemma 2.3. The constants δ>0 and D>0 exists such that
emtP(Z(t)=k|Z(0)=l)≤De−δll−12eke−mt,t>0,k≥1,l≥l0:=[1α]+1, |
where α is defined as in Lemma 2.2.
Proof. It can be obtained from (2.7) that
P(Z(t)=k|Z(0)=l)=ekh/emtGl(e−h/emt,t)P(Sl(h,t)=k), k≥1. | (2.23) |
Multiplying both sides of (2.23) by emt, specifying h=1, and using the definition of the generating function (2.1) together with Lemma 2.2, for l≥l0, we have
emtP(Z(t)=k|Z(0)=l)≤ekh/emtGl(e−1/emt,t)supt,kemtP(Sl(1,t)=k)≤ek/emt[F0l(e−1/emt,t)H(e−1/emt,t)]C(1)l−12≤C(1)(F0(e−1/emt,t))ll−12ek/emt, | (2.24) |
where C(1) is defined in Lemma 2.2. Note that
F0(e−1/emt,t)=E[e−Z0(t)/emt]t→∞⟶E[e−W]∈(0,1). |
There is a constant δ>0 such that supt>0F0(e−1/emt,t)≤e−δ<1, which is proved by substituting the inequality above into (2.24).
For sake of the subsequent discussions, we define the Laplace transform of the normalized random variables V, W, and I as follows:
ϕV(u):=E[e−uV],ϕW(u):=E[e−uW],ϕI(u):=E[e−uI],u≥0. | (2.25) |
Moreover, we have ϕV(u)=ϕW(u)ϕI(u), since Z0(t) and Y(t) are independent.
Proposition 2.3. (Iterative functional equations for the Laplace transform) If b0=0, then the following functional equations hold:
ϕV(emsu)=G(ϕW(u),s),u≥0,s≥0;ϕW(emsu)=F0(ϕW(u),s), u≥0,s≥0;ϕI(emsu)=H(ϕW(u),s), u≥0,s≥0. | (2.26) |
Proof. The conclusion can be obtained from (2.1) together with (2.25).
Record the characteristic function as ϕlW(a)ϕI(a): = (E[eiaW])lE[eiaI], where ϕlW(a)ϕI(a) is the characteristic function of w∗l∗i(a).
Lemma 2.4. For any x∈[a,b], there are constants C>0 and λ>0 such that
w∗l∗i(x)≤Cλl,l≥1. | (2.27) |
Proof. This follows from Lemma 5 in Athreya [4].
Lemma 2.5. Suppose that E[Z(1)logZ(1)]<+∞ and b1l+a0+m<0. Then for 0<a<b and l≥1,
limt→∞12π∫πemt−πemtGl(eix/emt,t)e−ixhdx=w∗l∗i(h) | (2.28) |
and this converges uniformly on [a,b].
Proof. This is achieved by the Fourier transform and decomposing the integral,
limt→∞12π∫πemt−πemtGl(eix/emt,t)e−ixhdx=limt→∞12π(∫−π−πemt+∫π−π+∫πemtπ)Gl(eix/emt,t)e−ixhdx:=H1+H2+H3. |
Applying the dominated convergence theorem (see [18], Theorem 16, p. 89), we have
H2=limt→∞12π∫π−πϕI(t)(x)(ϕW(t)(x))le−ixhdx=12π∫π−πϕI(x)(ϕW(x))le−ixhdx, |
where ϕI(t)(x):=E[eixI(t)] and ϕW(t)(x):=E[eixW(t)].
For
H3=limt→∞12π∫πemtπGl(eix/emt,t)e−ixhdx, |
the family of functions {eixξ;ξ≥1} is equicontinuous with respect to x and gives
F0(eixe−mt,t)t→∞⟶ϕW(x) and H(eixe−mt,t)t→∞⟶ϕI(x) |
uniformly with respect to x≥π. Hence, we can examine ∫πemtπGl(eix/emt)e−ixhdx by replacing ∫πemtπϕI(x)(ϕW(x))le−ixhdx.
Notice that |e−ixh|=1 and thus
12π∫πemtπϕI(x)(ϕW(x))le−ixhdx≤12π∫πemtπϕI(x)(ϕW(x))ldx=12π∫πemtπGl(ϕ(xe−mt),t)dx=12π∫ππe−mtemtGl(ϕ(u),t)du. | (2.29) |
The last equation is obtained by substituting x=uemt.
Assume that
Gl(ϕ(u),t)=∞∑j=0plj(t)ϕj(u)=Ql(ϕ(u),t)e(b1l+a0)t, |
where Ql(⋅,⋅) is as defined in (2.3). Then
H3≤limt→∞12π∫ππe−mtQl(ϕ(u),t)e(b1l+a0+m)tdu. |
Due to the assumption a0+b1l+m<0, it is obvious that e(a0+b1l+m)t→0 as t→∞. Moreover, the interval of integration is also finite. Combining this with the convergence of Ql(⋅,⋅), we have H3<∞. In the same way, we obtain H1<∞. Therefore
12π∫πemt−πemtGl(eix/emt,t)e−ixhdx |
converges to
limt→∞12π∫πemt−πemtϕI(x)(ϕW(x))le−ixhdx |
uniformly with respect to h. Finally, since ϕW(x) and ϕI(x) are the Fourier transformation of the probability density functions w(x) and i(x), combining these with the conclusion of Lemma 8 in Dubuc and Seneta [19] completes the conclusion.
Throughout this paper, we suppose that the generating function B(u) is aperiodic, i.e., the greatest common divisor of the set {i−j;i≠j,bibj>0,i⩾1,i⩾1} is 1. Assume that the sequence {kt} satisfies kt→∞ and kte−mt→0 as t→∞. Define s2:=logktm. Then we certainly have s2<t for a large enough t.
In the subsequent discussions, we always assume that the conditions (A1)−(A3) hold. On this basis, we give the main results of this paper.
Theorem 3.1. If we suppose that the generating function B(u) is aperiodic and the assumptions (A1)−(A3) hold, then
P(Z(t)=kt)=e−mtv(kt/emt)(1+o(1)), |
where the sequence {kt} satisfies kt→∞ and kt=o(emt) as t→∞.
Theorem 3.2. Suppose that the assumptions (A1)−(A3) hold and the generating function B(u) is aperiodic, then
P(0≤Z(t)≤kt)=FV(kt/emt)(1+o(1)), |
where the sequence {kt} are defined as in Theorem 3.1 and FV(x)=P(V≤x).
Depending on Lemma 2.5, it is easy to conclude the following local limit theorem.
Theorem 3.3. (Local limit theorem) Suppose that the assumptions (A1)−(A3) hold and the generating function B(u) is aperiodic if the integer sequence {kt} satisfies kt→∞, kt/emt→h, and h>0 is a constant, then
limt→∞emtP(Z(t)=kt|Z(0)=l)=w∗l∗i(h). |
In this section, we present detailed proofs related to the main results.
Proof. According to the Markov property
P(Z(t)=kt)=∞∑l=1P(Z(s1)=l)P(Z(t)=kt|Z(s1)=l)=∞∑l=1P(Z(s1)=l)P(Z(s2)=kt|Z(0)=l), | (4.1) |
where s1+s2=t. There is an integer N>1 such that
P(Z(t)=kt)=N−1∑l=1P(Z(s1)=l)P(Z(s2)=kt|Z(0)=l)+∞∑l=NP(Z(s1)=l)P(Z(s2)=kt|Z(0)=l):=I1(N,t)+I2(N,t). |
Next, we analyze the rate of convergence of I1(N,t) and I2(N,t). We begin with the second part I2(N,t). For a sufficiently large N such that N≥l0, by Lemma 2.3, there are D>0 and δ>0 such that
ems2∞∑l=NP(Z(s1)=l)P(Z(s2)=kt|Z(0)=l)≤D∞∑l=Ne−δll−1/2ekt/ems2P(Z(s1)=l)≤DN−1/2∞∑l=Ne−δlekt/ems2P(Z(s1)=l)≤DN−1/2G(e−δ,s1)ekt/ems2≤DN−1/2σs1. | (4.2) |
The last inequality holds by G(e−δ;s1)=Ee−Z(s1)δ<1≤Cσs1 together with the definition of σ, where s1 does not grow when t and C can be chosen appropriately.
The treatment of I1(N,t) is given below. On the one hand, according to Lemma 2.5, for l≥1,
limt→∞12π∫πemt−πemtGl(eiy/emt,t)e−iyxdy=w∗l∗i(x) | (4.3) |
holds uniformly on [e−m,1].
On the other hand, if ys2ems2→a as s2→∞, where a>0 is a constant, using the inversion formula
P(Z(s2)=ys2|Z(0)=l)=12π∫π−πGl(eiu,s2)e−iuys2du. |
Then kt=1 for all t. Setting u=v/ems2, we have
ems2P(Z(s2)=kt|Z(0)=l)=12π∫πems2−πems2Gl(eiv/ems2,s2)e−ivdv. | (4.4) |
According to (4.3)–(4.4) and Lemma 2.5
limt→∞[ems2P(Z(s2)=kt|Z(0)=l)−w∗l∗i(1)]=0. | (4.5) |
Hence
ems2N−1∑l=1P(Z(s1)=l)P(Z(s2)=kt|Z(0)=l)=[N−1∑l=1P(Z(s1)=l)w∗l∗i(1)](1+o(1)). | (4.6) |
Thus, for any fixed N
I1(N,t)=e−ms2N−1∑l=1P(Z(s1)=l)w∗l∗i(1)[1+o(1)]=e−ms2(∞∑l=1−∞∑N)P(Z(s1)=l)w∗l∗i(1)[1+o(1)]. | (4.7) |
With Lemma 2.4, there are C>0 and λ∈(0,1) such that w∗l∗i(1)≤Cλl for all l>1, and hence
e−ms2∞∑l=NP(Z(s1)=l)w∗l∗i(1)≤Ce−ms2∞∑l=NP(Z(s1)=l)λl. | (4.8) |
By (4.8) and the definition of G(⋅,⋅), and taking the constant λ1∈(λ,1), similarly to the proof of inequality (4.2), we have
∞∑l=NP(Z(s1)=l)w∗l∗i(1)≤C(λ/λ1)N∞∑l=NP(Z(s1)=l)λl1≤C(λ/λ1)NG(λ1,s1)≤Ce−δNσs1, | (4.9) |
where δ is a positive constant.
According to (4.1)–(4.2), (4.6), and (4.9)
P(Z(t)=kt)=e−ms2[∞∑l=1P(Z(s1)=l)w∗l∗i(1)][1+o(1)]+O(e−ms2N−1/2σs1). | (4.10) |
If we take the equality ϕ(uems1)=ϕW(uems1)lϕI(uems1)=G(ϕ(u),s1) in the form of a density function, then for any x≥0
∞∑l=1P(Z(s1)=l)w∗l∗i(x)=v(x/ems1)/ems1=v(xkt/emt)/em(t−s2). | (4.11) |
By setting x=kt/ems2=1 in the equality above, then (4.10) becomes
P(Z(t)=kt)=e−mtv(kt/emt)[1+o(1)]+O(e−ms2N−1/2σs1). | (4.12) |
Let t go to infinity first and then let N go to infinity in the equality above. The proof is completed.
Proof. By the Markov property, we have
P(Z(t)≤kt)=∞∑l=1P(Z(s1)=l)P(Z(s2)≤kt|Z(0)=l). | (4.13) |
According to the branching property combined with the independence between immigration and branching, it follows that
P(Z(s2)≤kt|Z(0)=l)=P(Z0(s2)+Y(s2)≤kt|Z(0)=l)≤P(Z0(s2)≤kt|Z(0)=l)P(Y(s2)≤kt)≤[P(Z0(s2)≤kt)]l. | (4.14) |
By (1.4)
P(Z0(s2)≤kt)=P(Z0(s2)ems2≤1)→∫10w(x)dx, t→∞, |
where w(x) is continuous in (0,∞). Thus, a constant η∈(0,1) and a sufficiently large t exist such that P(Z0(s2)≤kt)≤η. By formula (4.14) and a sufficiently large N, we have C and δ>0
∞∑l=NP(Z(s1)=l)P(Z(s2)≤kt|Z(0)=l)≤∞∑l=NP(Z(s1)=l)ηl≤Cσs1e−δN. | (4.15) |
If we assume FW(x)=P(W≤x) and FI(x)=P(I≤x), then by (1.3), together with the continuity of the distribution function
P(Z(s2)≤xems2|Z(0)=l)→F∗lW∗FI(x) |
uniformly in x>0. Hence, we have
limt→∞supk≥1|P(Z(s2)≤kt|Z(0)=l)−F∗lW∗FI(1)|=0. | (4.16) |
According to (4.13), (4.15), and (4.16)
P(Z(t)≤kt)=[∞∑l=1P(Z(s1)=l)F∗lW∗FI(1)](1+o(1))+O(σs1e−δN). | (4.17) |
Moreover, by the definition of s2 and the inequality F∗lW∗FI(1)≥F∗lW∗FI(1/em), there is a constant C∈[0,1] such that
∞∑l=1P(Z(s1)=l)F∗lW∗FI(1)≥P(Z(s1)=1)F∗lW∗FI(1/em)≥Cσs1. |
Hence, (4.17) can be simplified to
P(Z(t)≤kt)=[∞∑l=1P(Z(s1)=l)F∗lW∗FI(1)][1+o(1)+O(e−δN)]. | (4.18) |
Integrating both sides of the density function (4.11), we have
FV(a/emk)=∞∑l=1P(Z(k)=l)F∗lW∗FI(a). | (4.19) |
Taking k=s1 and a=1 in equation above and substituting this into (4.18)
P(Z(t)≤kt)=FV(kt/emt)(1+o(1)+O(e−δN)). | (4.20) |
Letting t→∞ and N→∞ completes this proof.
Proof. This is similar to the proofs in [20] (Theorem 7.1, p. 105)
kte−mt→h(t→∞), |
then by the inversion formula
P(Z(t)=kt|Z(0)=l)=12π∫π−πGl(eiu,t)e−iuktdu. |
If we set u=ve−mt, then
emtP(Z(t)=kt|Z(0)=l)=12π∫πemt−πemtGl(eiv/emt,t)e−ivkt/emtdv. |
Therefore
limt→∞[emtP(Z(t)=kt|Z(0)=l)]=limt→∞12π∫πemt−πemtGl(eiv/emt,t)e−ivkt/emtdv=w∗l∗i(h). |
The proof is completed.
In this paper, we discuss a continuous-time supercritical branching process with immigration (MBPI). We mainly research the local lower deviation probabilities and the global lower deviation probabilities, obtain some related results such as local limit theorem and some related estimates of the MBPIs, which generalized the results of discrete-time branching processes to continuous-time cases.
Juan Wang: Conceptualization, writing—review and editing, funding acquisition; Chao Peng: Conceptualization, writing—original draft preparation. All authors have read and approved the final version of the manuscript for publication.
The authors declare that they have not used artificial intelligence (AI) tools in the creation of this article.
This work is substantially supported by the National Natural Sciences Foundation of China (No.11901392).
The authors declare no conflicts of interest.
[1] | K. B. Athreya, P. E. Ney, Branching processes, Berlin, Heidelberg: Springer, 1972. https://doi.org/10.1007/978-3-642-65371-1 |
[2] |
J. Li, L. Cheng, L. Li, Long time behaviour for Markovian branching-immigration systems, Discrete Event Dyn. Syst., 31 (2021), 37–57. https://doi.org/10.1007/s10626-020-00323-z doi: 10.1007/s10626-020-00323-z
![]() |
[3] |
K. B. Athreya, P. Ney, The local limit theorem and some related aspects of super-critical branching processes, Trans. Am. Math. Soc., 152 (1970), 233–251. https://doi.org/10.1090/S0002-9947-1970-0268971-X doi: 10.1090/S0002-9947-1970-0268971-X
![]() |
[4] |
K. B. Athreya, Large deviation rates for branching processes–Ⅰ. Single type case, Ann. Appl. Probab., 4 (1994), 779–790. https://doi.org/10.1214/aoap/1177004971 doi: 10.1214/aoap/1177004971
![]() |
[5] |
P. E. Ney, A. N. Vidyashankar, Harmonic moments and large deviation rates for supercritical branching processes, Ann. Appl. Probab., 13 (2003), 475–489. https://doi.org/10.2307/1193154 doi: 10.2307/1193154
![]() |
[6] |
P. E. Ney, A. N. Vidyashankar, Local limit theory and large deviations for supercritical branching processes, Ann. Appl. Probab., 14 (2004), 1135–1166. https://doi.org/10.1214/105051604000000242 doi: 10.1214/105051604000000242
![]() |
[7] |
K. Fleischmann, V. Wachtel, Lower deviation probabilities for supercritical Galton-Watson processes, Ann. Inst. H. Poincaré Probab. Statist., 43 (2007), 233–255. https://doi.org/10.1016/j.anihpb.2006.03.001 doi: 10.1016/j.anihpb.2006.03.001
![]() |
[8] |
E. Seneta, On the supercritical Galton-Watson process with immigration, Math. Biosci., 7 (1970), 9–14. https://doi.org/10.1016/0025-5564(70)90038-6 doi: 10.1016/0025-5564(70)90038-6
![]() |
[9] |
A. Pakes, On supercritical galton-watson processes allowing immigration, J. Appl. Probab., 11 (1974), 814–817. https://doi.org/10.2307/3212564 doi: 10.2307/3212564
![]() |
[10] |
W. Chu, W. V. Li, Y. X. Ren, Small value probabilities for supercritical branching processes with immigration, Bernoulli, 20 (2014), 377–393. https://doi.org/10.3150/12-BEJ490 doi: 10.3150/12-BEJ490
![]() |
[11] |
J. Liu, M. Zhang, Large deviation for supercritical Galton-Watson processes with immigration, Acta. Math. Sin., 32 (2016), 893–900. https://doi.org/10.1007/s10114-016-5437-z doi: 10.1007/s10114-016-5437-z
![]() |
[12] |
Q. Sun, M. Zhang, Harmonic moments and large deviations for supercritical branching processes with immigration, Front. Math. China, 12 (2017), 1201–1220. https://doi.org/10.1007/s11464-017-0642-3 doi: 10.1007/s11464-017-0642-3
![]() |
[13] |
Q. Sun, M. Zhang, Lower deviations for supercritical branching processes with immigration, Front. Math. China, 16 (2021), 567–594. https://doi.org/10.1007/s11464-021-0922-9 doi: 10.1007/s11464-021-0922-9
![]() |
[14] |
D. Li, M. Zhang, Harmonic moments and large deviations for a critical Galton-Watson processes with immigration, Sci. China. Math., 64 (2021), 1885–1904. https://doi.org/10.1007/s11425-019-1676-x doi: 10.1007/s11425-019-1676-x
![]() |
[15] |
J. Li, L. Cheng, A. G. Pakes, A. Chen, L. Li, Large deviation rates for Markov branching processes, Anal. Appl., 18 (2020), 447–468. https://doi.org/10.1142/S0219530519500209 doi: 10.1142/S0219530519500209
![]() |
[16] |
J. Li, A. Chen, A. G. Pakes, Asymptotic properties of the Markov branching process with immigration, J. Theor. Probab., 25 (2012), 122–143. https://doi.org/10.1007/s10959-010-0301-z doi: 10.1007/s10959-010-0301-z
![]() |
[17] | V. V. Petrov, Sums of independent random variables, Berlin, Heidelberg: Springer, 1975. https://doi.org/10.1007/978-3-642-65809-9 |
[18] | H. L. Royden, Real analysis, New York: Macmillan, 1968. |
[19] |
S. Dubuc, E. Seneta, The local limit theorem for the galton-watson process, Ann. Probab., 4 (1976), 490–496. https://doi.org/10.1214/aop/1176996100 doi: 10.1214/aop/1176996100
![]() |
[20] | S. Asmussen, H. Hering, Branching processes, Boston, MA: Birkhäuser, 1983. https://doi.org/10.1007/978-1-4615-8155-0 |