Processing math: 100%
Research article Special Issues

Near-resonances and detuning in classical and quantum mechanics

  • From the point of view of perturbation theory, (perturbations of) near-resonant systems are plagued by small denominators. These do not affect (perturbations of) fully resonant systems; so it is in many ways convenient to approximate near resonant systems as fully resonant ones, which correspond to considering the "detuning" as a perturbation itself. This approach has proven very fruitful in Classical Mechanics, but it is also standard in (perturbations of) Quantum Mechanical systems. Actually, its origin may be traced back (at least) to the Rayleigh-Ritz method for computing eigenvalues and eigenvectors of perturbed matrix problems. We will discuss relations between these approaches, and consider some case study models in the different contexts.

    Citation: G. Gaeta, G. Pucacco. Near-resonances and detuning in classical and quantum mechanics[J]. Mathematics in Engineering, 2023, 5(1): 1-44. doi: 10.3934/mine.2023005

    Related Papers:

    [1] Virginia Giorno, Serena Spina . On the return process with refractoriness for a non-homogeneous Ornstein-Uhlenbeck neuronal model. Mathematical Biosciences and Engineering, 2014, 11(2): 285-302. doi: 10.3934/mbe.2014.11.285
    [2] Giuseppe D'Onofrio, Enrica Pirozzi . Successive spike times predicted by a stochastic neuronal model with a variable input signal. Mathematical Biosciences and Engineering, 2016, 13(3): 495-507. doi: 10.3934/mbe.2016003
    [3] Aniello Buonocore, Luigia Caputo, Enrica Pirozzi, Maria Francesca Carfora . A simple algorithm to generate firing times for leaky integrate-and-fire neuronal model. Mathematical Biosciences and Engineering, 2014, 11(1): 1-10. doi: 10.3934/mbe.2014.11.1
    [4] Meng Gao, Xiaohui Ai . A stochastic Gilpin-Ayala mutualism model driven by mean-reverting OU process with Lévy jumps. Mathematical Biosciences and Engineering, 2024, 21(3): 4117-4141. doi: 10.3934/mbe.2024182
    [5] Yongxin Gao, Shuyuan Yao . Persistence and extinction of a modified Leslie-Gower Holling-type Ⅱ predator-prey stochastic model in polluted environments with impulsive toxicant input. Mathematical Biosciences and Engineering, 2021, 18(4): 4894-4918. doi: 10.3934/mbe.2021249
    [6] Aniello Buonocore, Luigia Caputo, Enrica Pirozzi, Maria Francesca Carfora . Gauss-diffusion processes for modeling the dynamics of a couple of interacting neurons. Mathematical Biosciences and Engineering, 2014, 11(2): 189-201. doi: 10.3934/mbe.2014.11.189
    [7] Virginia Giorno, Amelia G. Nobile . Exact solutions and asymptotic behaviors for the reflected Wiener, Ornstein-Uhlenbeck and Feller diffusion processes. Mathematical Biosciences and Engineering, 2023, 20(8): 13602-13637. doi: 10.3934/mbe.2023607
    [8] Giacomo Ascione, Enrica Pirozzi . On a stochastic neuronal model integrating correlated inputs. Mathematical Biosciences and Engineering, 2019, 16(5): 5206-5225. doi: 10.3934/mbe.2019260
    [9] Buyu Wen, Bing Liu, Qianqian Cui . Analysis of a stochastic SIB cholera model with saturation recovery rate and Ornstein-Uhlenbeck process. Mathematical Biosciences and Engineering, 2023, 20(7): 11644-11655. doi: 10.3934/mbe.2023517
    [10] Huili Wei, Wenhe Li . Dynamical behaviors of a Lotka-Volterra competition system with the Ornstein-Uhlenbeck process. Mathematical Biosciences and Engineering, 2023, 20(5): 7882-7904. doi: 10.3934/mbe.2023341
  • From the point of view of perturbation theory, (perturbations of) near-resonant systems are plagued by small denominators. These do not affect (perturbations of) fully resonant systems; so it is in many ways convenient to approximate near resonant systems as fully resonant ones, which correspond to considering the "detuning" as a perturbation itself. This approach has proven very fruitful in Classical Mechanics, but it is also standard in (perturbations of) Quantum Mechanical systems. Actually, its origin may be traced back (at least) to the Rayleigh-Ritz method for computing eigenvalues and eigenvectors of perturbed matrix problems. We will discuss relations between these approaches, and consider some case study models in the different contexts.



    Stochastic age-dependent population system has an increasingly important status in biomathematics, being the main research direction in biology and ecology in recent years. Especially, one of the most striking and meaningful problems in the study of stochastic age-dependent population system is its numerical scheme because of its nonlinear structure and non-existent explicit solutions. In [1], Li et al. first introduced Poisson jumps into stochastic age-dependent population system and proposed the following model:

    {dtPt=[Ptaμ(t,a)Pt+f(t,Pt)]dt+g1(t,Pt)dBt+h(t,Pt)dNt     (t,a)Q,P(0,a)=P0(a),   a[0,A],P(t,0)=A0β(t,a)P(t,a)da,    t[0,T], (1.1)

    where T>0, A is the maximal age of the population species and A>0, Q=(0,T)×(0,A). dtPt is the differential of Pt relative to t, i.e., dtPt=Pttdt. Pt=P(t,a) denotes the population density of age a at time t, μ(t,a) denotes the mortality rate of age a at time t, β(t,a) denotes the fertility rate of females of a at time t, f(t,Pt) denotes effects of external environment for population system, g1(t,Pt) is a diffusion coefficient, h(t,Pt) is a jump coefficient (it represents the size of the population systems increases or decreases drastically because brusque variations from earthquakes, floods, immigrants and so on), Bt is a Brownian motion, Nt is a Poisson process with intensity λ>0. Then, they investigated the convergence of Euler method for model (1.1). Since then, an increasing number of authors have analyzed stochastic age-dependent population models with Poisson jumps, and many significant results have been obtained (see e.g., [2,3,4,5,6,7,8,9,10,11]). For example, Wang and Wang [3] established the semi-implicit Euler method for stochastic age-dependent population models with Poisson jumps and discussed the convergence order of numerical solutions. Tan et al. [5] presented a split-step θ (SSθ) method of stochastic age-dependent population models with Poisson jumps, and the exponential stability of the model was established. Pei et al. [8] constructed two types of numerical methods for stochastic age-dependent population models with Poisson jumps, which are compensated and non-compensated. Then the asymptotic mean-square boundedness is discussed for numerical scheme.

    In the above-mentioned model (1.1), we can easily see that the effects of randomly environmental variations of parameter μ are described as a linear function of Gaussian white noise [12,13], that is μdtμdt+σdBt (where σ2 represents the intensity of Bt). Obviously, it is unreasonable to use linear function of Gaussian white noise to simulate parameters perturbation in a randomly varying environment. In [14], Duffie proposed to use a mean-reverting Ornstein-Uhlenbeck process to describe parameters which fluctuate around an average value. Up until now, the mean-reverting OU process have been extensively discussed in [15,16,17,18]. Zhao et al. [16] analyzed stationary distribution for stochastic competitive model incorporating the OU process. Wang et al. [17] introduced the OU process into the Susceptible-Infected-Susceptible (SIS) epidemic model and investigated its threshold. However, to the authors' knowledge, there is no literature to consider the OU process into the stochastic age-dependent population system with Poisson jumps. On the other hand, we find that the influence of time delay are not considered in the above papers [1,2,3,4,5,6,7,8,9,11]. There are very few papers in the literature that take time delay into account in the stochastic age-dependent population system so far (see e.g., [19,20]). Therefore, based on the above analysis, studying stochastic age-dependent delay population jumps equations, coupled with mean-reverting OU process have more practical significance.

    Obviously, the age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps have no explicit solution. Thus, numerical approximation schemes should be developed as a essential and powerful tool to explore its properties. The numerical schemes of stochastic age-dependent population system have been extensively researched by many scholars, for example, [1,3,4,5,8,9,19,21,22,23]. However, due to the fact that the coefficients f, g1 and h of model (1.1) are particularly complex functions, so using the existing numerical methods to approximate the age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps will result in slow convergence and very high computational cost. Recently, Jankovic and Ilic [24] introduced a Taylor approximation method for stochastic differential equations and proved that its convergence rate and computational cost is better than other numerical methods such as the Euler method, the semi-implicit Euler method and SSθ method mentioned in [1,2,3,4,5,6,7,8,9,11]. Motivated by Jankovic et al., we construct a Taylor approximation scheme for the age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps in this paper. Furthermore, the convergence between the exact solutions and numerical solutions is investigated.

    The highlights of the present paper are summarized as follows:

    ● Age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps are given.

    ● To improve the convergence speed and reduce cost, the Taylor approximation scheme for the age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps is developed.

    ● The convergence and convergence order of Taylor approximation scheme are discussed.

    The arrangement of this paper is as follows. In section 2, we establish the age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps, as well as introduce some notations and preliminaries. Then, the pth moments boundedness of exact solutions for age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps are presented. In section 3, we propose a Taylor approximation scheme for a age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps, and the convergence theory for the numerical method is proved. In section 4, we present some numerical simulations to demonstrate our theoretical results. Section 5 presents the conclusions of our research.

    In the above model (1.1), the authors took advantage of a traditional parameter perturbation method to reflect the effect of environmental noise, (μ(t,a)Pt+f(t,Pt))dt when it is stochastically perturbed with (μ(t,a)Pt+f(t,Pt))dt+g1(t,Pt)dBt. It is worth mentioning that Cai et al. [18] pointed out that due to environmental continuous fluctuations, the mortality rate μ(t,a) can not be described by a linear function of Gaussian white noise. In order to model the randomly varying environmental fluctuations in μ(t,a), we introduce the following mean-reverting Ornstein-Uhlenbeck process for μ(t,a) inspired by [16]:

    {μ(t,a)=μ1(t)μ2(a)dμ1(t)=η(μeμ1(t))dt+ξdBt (2.1)

    where we assume that μ1(t) represents the mortality rate at time t and μ2(a) represents the mortality rate at age a. All parameters η, μe and ξ are positive constants. η is the reversion rate, μe is the mean reversion level or long-run equilibrium of growth rate μ1(t), ξ is the intensity of volatility.

    For (2.1), applying the stochastic integral format, we obtain the explicit form of the solution as:

    μ1(t)=μe+(μ01μe)eηt+ξt0eη(ts)dBt (2.2)

    where (μ01:=μ1(0)). It is not difficult to see that the expected value of μ1(t) is

    E[μ1(t)]=μe+(μ01μe)eηt (2.3)

    and variance value of μ1(t) is

    Var[μ1(t)]=ξ22η(1e2ηt). (2.4)

    Combining (2.2), (2.3) and (2.4), we easily have that the term ξt0eη(ts)dBt satisfies the normal distribution E(0,ξ22η(1e2ηt)). Then we obtain ξt0eη(ts)dB(t) being equal to ξ2η1e2ηtdBtdt a.e..

    Therefore, we can rewrite (2.2) in the following form [17,18]

    μ1(t)=μe+(μ01μe)eηt+σ(t)dBtdt, (2.5)

    where σ(t)=ξ2η1e2ηt. A conceptual problem immediately occurs in that dBtdt is not defined except in a generalized sense.

    Replacing μ(t,a) in model (1.1) with (2.2) and (2.5) and rearranging leads to the following stochastic age-dependent population equation:

    {dtPt=[Pta(μe+(μ01μe)eηt)μ2(a)Pt+f(t,Pt)]dt+g(t,Pt)dBt+h(t,Pt)dNt      (t,a)(0,T)×(0,A),P(0,a)=P0(a), a[0,A],P(t,0)=A0β(t,a)P(t,a)da,  t[0,T], (2.6)

    where g(t,Pt)dBt contains g1(t,Pt)dBt and σ(t)μ2(a)PtdBt.

    On the other hand, due to the time delay is unavoidable in a real world. Motivated by [19] and [25], we derive the following system:

    {dtPt=[Pta(μe+(μ01μe)eηt)μ2(a)Pt+f(t,Pt,Ptτ)]dt+g(t,Pt,Ptτ)dBt+h(t,Pt,Ptτ)dNt,         (t,a)(0,T)×(0,A),P(t,a)=ϕ(t,a), (t,a)[τ,0]×[0,A],P(t,0)=A0β(t,a)P(t,a)da,   t[0,T], (2.7)

    where Pt=P(t,a) denotes the population density of age a at time t, Ptτ=P(tτ,a) denotes the population density of age a at time tτ, τ is time delay and τ>0. f(t,Pt,Ptτ) denotes the effects of external environment for the population system, g(t,Pt,Ptτ) is a diffusion coefficient, h(t,Pt,Ptτ) is a jump coefficient, ϕt=ϕ(t,a) denotes the histories of the population density of age a at time t, β(t, a) denotes the fertility rate of females of age a at time t. A is the maximal age of the population species, so P(t,a)=0,aA. Nt is a Poisson process with intensity λ>0. The explanation of the other symbols was given under Eq (2.1). In the following section, we concentrate on studying model (2.7).

    Let V=D1([0,A]){φ|φL2([0,A]),where φa represent the generalized partial derivatives}, V be a Sobolev space. D=L2[0,A] such that VDDV. V is the dual space of V. We denote by , and the norms in V,D and V respectively; by , the duality product between V and V, and by (,) the scalar product in D.

    For simplicity, we introduce some notations. Throughout this paper, unless otherwise, let (Ω,F,P) be a complete probability space with filtration {Ft}t0 satisfying the usual conditions (i.e. it is increasing and right continuous while F0 contains all P-null sets), and let E signify the expectation corresponding to P. For an operator HL(M,D) on the space of all bounded linear operators from M into D, we denote by H2 the Hilbert-Schmidt norm, i.e., B2=tr(HWH)T.

    Let τ>0 and C=C([τ,0];D) be the space of all continuous function from [0, T] into H with sup-norm ψC=supτs0|ψ(s)|, LPV=LP([0,T];V) and LPD=LP([0,T];D). Moreover, let F0-measurable, CbF0([τ,0];D) denote the family of all almost surely bounded, F0-measurable C=C([τ,0];D) -value random variables. For a pair of real numbers a and b, we use ab=max(a,b). If G is a set, its indicator function by 1G, namely 1G(x) = 1 if xG and 0 otherwise.

    The integer version of Eq (2.7) is given by

    Pt=P0t0Psadst0[μe+(μ01μe)eηs)]μ2(a)Psds+t0f(s,Ps,Psτ)ds+t0g(s,Ps,Psτ)dBs+t0h(s,Ps,Psτ)dNs. (2.8)

    For the existence and uniqueness of the solution, we assume that the following assumptions are satisfied:

    (A1) μ(t,a) and β(t,a) are nonnegative measurable, such that

    {0μ_μ2(a)<ˉμ<,0β(t,a)ˉβ<.

    (A2) f(t,0,0)=g(t,0,0)=h(t,0,0)=0.

    (A3) The Lipschitz and linear growth conditions: there exists a positive constant K such that

    |f(t,x1,y1)f(t,x2,y2)|g(t,x1,y1)g(t,x2,y2)2|h(t,x1,y1)h(t,x2,y2)|

    K(x1x2C+y1y2C),

    |f(t,x,y)|2g(t,x,y)22|h(t,x,y)|22K2(x2C+y2C)

    for x,y,x1,x2,y1,y2C.

    (A4) There exists constants ˜K,ˉK>0 and γ(0,1] such that for τs0, 0aA and r2

    E|ϕtϕs|r˜K(ts)γ.

    Consequently,

    E|ϕt|r<.

    (A5) f, g and h have Taylor approximations in the second argument, up to α1th, α2th and α3th derivatives, denoted as f(α1)x(t,x,y), g(α2)x(t,x,y) and h(α3)x(t,x,y), respectively.

    (A6) f(α1+1)Pt(t,Pt,Ptτ), g(α2+1)Pt(t,Pt,Ptτ) and h(α3+1)Pt(t,Pt,Ptτ) are uniformly bounded, i.e. there exist positive constants K1, K2 and K3 satisfying

    {sup[0,T]×[0,A]|f(α1+1)Pt(t,Pt,Ptτ)|K1,sup[0,T]×[0,A]|g(α2+1)Pt(t,Pt,Ptτ)|K2,sup[0,T]×[0,A]|h(α3+1)Pt(t,Pt,Ptτ)|K3.

    Throughout the following analysis, for the purpose of simplicity, we will use C,C1,C2, to stand for generic constants that depend upon K and T, but not upon Δ. The precise value of these constants may be determined via the proof. Our first theorem shows the existence and uniqueness of the strong solution for the model (2.7).

    Theorem 2.1. Under the assumptions (A1)(A4), for t[0,T], Eq (2.7) has a unique strong solution.

    Proof. The proof of this theorem is standard (see Zhang et al. [26]) and hence is omitted.

    Moreover, the pth moment boundedness of the true solution Pt of the model (2.7) is proved in the following theorem.

    Theorem 2.2. Under the assumptions (A1)(A4), for each q2, there exists a constant C such that

    E[supτtT|Pt|q]C. (2.9)

    Proof. Form (2.8), applying Itô's formula [25] to |Pt|q yields

    |Pt|q= |P0|q+t0q|Ps|q2Psa(μe+(μ01μe)eηs)μ2(a)Ps,Psds+t0q|Ps|q2(f(s,Ps,Psτ),Ps)ds+t0q(q1)2|Ps|q2g(s,Ps,Psτ)22ds+t0q|Ps|q2(Ps,g(s,Ps,Psτ))dBs+t0q|Ps|q2(Ps,h(s,Ps,Psτ))dNs+t0q(q1)2|Ps|q2h(s,Ps,Psτ)22dNs= |P0|q+t0q|Ps|q2Psa(μe+(μ01μe)eηs)μ2(a)Ps,Psds+t0q|Ps|q2(f(s,Ps,Psτ),Ps)ds+t0q(q1)2|Ps|q2g(s,Ps,Psτ)22ds+t0q|Ps|q2(Ps,g(s,Ps,Psτ))dBs+t0q|Ps|q2(Ps,h(s,Ps,Psτ))dˉNs+λt0q|Ps|q2(Ps,h(s,Ps,Psτ))ds+t0q(q1)2|Ps|q2|h(s,Ps,Psτ)|2dˉNs+λt0q(q1)2|Ps|q2|h(s,Ps,Psτ)|2ds, (2.10)

    where ˉNs=Nsλs is a compensated Poisson process.

    Since

    Psa,Ps=A0Psda(Ps)= 12(A0β(t,a)Psda)2 12A0β2(t,a)daA0P2sda 12ˉβ2A2|Ps|2, (2.11)

    by the assumptions (A1)-(A3), we get that

    |Pt|q |P0|q+q(ˉβ2A22+μ01ˉμ)t0|Ps|qds+q(q1)K2t0|Ps|q2(Ps2C+Psτ2C)ds+Kqt0|Ps|q1(PsC+PsτC)ds+qt0|Ps|q2(Ps,g(s,Ps,Psτ))dBs+qt0|Ps|q2(Ps,h(s,Ps,Psτ))dˉNs+q(q1)2t0|Ps|q2|h(s,Ps,Psτ)|2dˉNs+λqKt0|Ps|q1(PsC+PsτC)ds+λq(q1)K2t0|Ps|q2(Ps2C+Psτ2C)ds |P0|q+q(ˉβ2A22+μ01ˉμ)t0|Ps|qds+2q(q1)K2t0supτus|Pu|qds+2Kqt0supτus|Pu|qds+qt0|Ps|q2(Ps,g(s,Ps,Psτ))dBs+qt0|Ps|q2(Ps,h(s,Ps,Psτ))dˉNs+2Kqλt0supτus|Pu|qds+q(q1)2t0|Ps|q2|h(s,Ps,Psτ)|2dˉNs+2λq(q1)K2t0supτus|Pu|qds |P0|q+q[(ˉβ2A22+μ01ˉμ)+2K+2(q1)K2+2Kλ+2λK2(q1)]t0supτus|Pu|qds+qt0|Ps|q2(Ps,g(s,Ps,Psτ))dBs+qt0|Ps|q2(Ps,h(s,Ps,Psτ))dˉNs+q(q1)2t0|Ps|q2|h(s,Ps,Psτ)|2dˉNs. (2.12)

    Note that for any t[0,T],

    E[supτut|Pu|q]=E[supτu0|Pu|q]E[sup0ut|Pu|q]. (2.13)

    Hence, we have

    E[supτut|Pu|q] E[supτu0|ϕu|q]+C1t0E[supτus|Pu|q]ds+qE[sup0stt0|Ps|q2(Ps,g(s,Ps,Psτ))dBs]+qE[sup0stt0|Ps|q2(Ps,h(s,Ps,Psτ))dˉNs]+q(q1)2E[sup0stt0|Ps|q2|h(s,Ps,Psτ)|2dˉNs], (2.14)

    where C1=q[(ˉβ2A22+μ01ˉμ)+2K+2(q1)K2+2Kλ+2λK2(q1)].

    Using the Burkholder-Davis-Gundy's inequality, we derive that

    E[sup0stt0|Ps|q2(Ps,g(s,Ps,Psτ))dBs]=E[sup0stt0|Ps|q2(Pq22s,g(s,Ps,Psτ))dBs]E[supτut|Pu|q2(t0(Pq22s,g(s,Ps,Psτ))dBs)]3E[supτut|Pu|q2(t0|Ps|q2g(s,Ps,Psτ)22ds)12]16qE[supτut|Pu|q]+C2E(t0|Ps|q2g(s,Ps,Psτ)22ds)16qE[supτut|Pu|q]+4C2K2(t0Esupτus|Pu|qds). (2.15)

    Similarly, we can obtain that

    E[sup0stt0|Ps|q2(Ps,h(s,Ps,Psτ))dˉNs]16qE[supτut|Pu|q]+4C3K2(t0Esupτus|Pu|qds) (2.16)

    and

    E[sup0stt0|Ps|q2|h(s,Ps,Psτ)|22dˉNs]13q(q1)E[supτut|Pu|q]+16C4K4(t0Esupτus|Pu|qds). (2.17)

    Substituting (2.15), (2.16) and (2.17) into (2.14) yields

    E[supτut|Pu|q] E[supτu0|ϕu|q]+12E[supτut|Pu|q]+(C1+4qC2K2+4qC3K2+8q(q1)C4K4)(t0Esupτus|Pu|qds). (2.18)

    Thus, the well-known Gronwall inequality obviously implies the desired equality (2.9).

    In this section, we will establish the Taylor approximation scheme for the stochastic age-dependent population Eq (2.7) and investigate the strong convergence between the true solutions and the numerical solutions derived from the Taylor approximation scheme.

    Let τj denote the jth jump of Ns occurrence time. For example, assume that jumps arrive at distinct, ordered times τ1<τ2<, let t1,t2,,tm be the deterministic grid points of [0,T]. We establish approximate solutions to (2.7) at a discrete set of times {τj}(j=1,2,). This set is the superposition of the random jump times of the Poisson process on [0,T] and a deterministic grid t1,t2,,tm and satisfy max{|τi+1τi|}<Δt (for the sake of simplicity, we denote Δt as Δ). It is quite clear that the random Poisson jump times can be computed without any knowledge of the realized path of (2.7).

    Next, we propose a Taylor approximation of the solutions of Eq (2.7). Without loss of any generality, given a step size Δ(0,1), we let tk=kΔ for k=0,1,2,3,,[τΔ], here [τΔ] is the integer part of τΔ. The continuous time Taylor approximate solution Qt=Q(t,a) to the stochastic age-dependent population Eq (2.7) can be defined by setting Q0=P0(a)=ϕ(0,a) and Q(t,0)=A0β(t,a)Qtda and forming

    {Qt=Q0t0Qsadst0(μe+(μ01μe)eηs)μ2(a)Qsds+t0α1j=0f(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)jds+t0α2j=0g(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)jdBs+t0α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)jdNs,     0tT,Qt=ϕ(t,a), τt0, (3.1)

    where Z1t=Z1(t,a)=[τΔ]k=0Qtk1[tk,tk+1)(t) and Z2t=Z2(t,a)=[τΔ]k=0Qtkτ1[tk,tk+1)(t) are step processes. That is, Z1t=Qtk and Z2t=Qtkτ for t[tk,tk+1) when k = 0, 1, 2, 3, , [τΔ].

    In this subsection, let us investigate the convergence of the Taylor approximate solutions of the stochastic age-dependent population Eq (2.7).

    In the following three lemmas we will show that Qt and Qtτ are close to Z1t and Z2t based on Lr, respectively.

    Lemma 3.1. For any q2, there exists a positive constant K4 such that

    Esupt[τ,T]|Qt|(α+1)2qK4, (3.2)

    where α=max{α1,α2,α3}.

    Proof. This proof is completed in Appendix A.

    Remark 3.1. If α1=α2=α3=0, then (A3) shows that the Taylor approximation solutions Qt admit finite moments (see, [27,28]).

    Lemma 3.2. Under the assumptions (A1),(A3),(A5), (A6), Lemma 3.1 and E|Qsa|r<K5 hold, K5 is a positive constant. For 2r(α+1)q, we have

    E|QtZ1t|rCΔr2. (3.3)

    Proof. For any t0, there exists an integer k0 such that t[tk,tk+1), we have

    QtZ1t=QtQtk=ttkQsadsttk(μe+(μ01μe)eηs)μ2(a)Qsds+ttkX1(s,Qs,Z1s,Z2s)ds+ttkX2(s,Qs,Z1s,Z2s)dBs+ttkX3(s,Qs,Z1s,Z2s)dNs,

    where

    X1(t,Qt,Z1t,Z2t)=α1j=0f(j)Z1t(t,Z1t,Z2t)j!(QtZ1t)j,X2(t,Qt,Z1t,Z2t)=α2j=0g(j)Z1t(t,Z1t,Z2t)j!(QtZ1t)j,X3(t,Qt,Z1t,Z2t)=α3j=0h(j)Z1t(t,Z1t,Z2t)j!(QtZ1t)j.

    By the elementary inequality, we further have

    E|QtZt|r 5r1[E|ttkQsads|r+E|ttk(μe+(μ01μe)eηs)μ2(a)Qsds|r+E|ttkX1(s,Qs,Z1s,Z2s)|r+E|ttkX2(s,Qs,Z1s,Z2s)dBs|r+E|ttkX3(s,Qs,Z1s,Z2s)dNs|r].

    Applying the Hölder inequality and moment inequality, we obtain

    E|QtZt|r 5r1[Δr1ttkE|Qsa|rds+(μ01ˉμ)rΔr1ttkE|Qs|rds+Δr1ttkE|X1(s,Qs,Z1s,Z2s)|rds+C1Δr21ttkEX2(s,Qs,Z1s,Z2s)r2dBs+E|ttkX3(s,Qs,Z1s,Z2s)dNs|r]. (3.4)

    For the jump integer, by virtue of the elementary inequality and Doob's inequality, we derive

    E|ttkX3(s,Qs,Z1s,Z2s)dNs|r= E|ttkX3(s,Qs,Z1s,Z2s)dˉNs+λttkX3(s,Qs,Z1s,Z2s)ds|r 2r1E|ttkX3(s,Qs,Z1s,Z2s)dˉNs|r+2r1E|λttkX3(s,Qs,Z1s,Z2s)ds|r C22r1Δr21ttkE|X3(s,Qs,Z1s,Z2s)|rds+2r1λrΔr1ttkE|X3(s,Qs,Z1s,Z2s)|rds. (3.5)

    Moreover, by the well-known mean value theorem, we observe that there exists a θ(0,1) such that

    ttkE|X1(s,Qs,Z1s,Z2s)|rds= ttkE|f(s,Qs,Z2s)[f(s,Qs,Z2s)X1(s,Qs,Z1s,Z2s)]|rds= ttkE|f(s,Qs,Z2s)f(α1+1)P(s,Z1s+θ(QsZ1s),Z2s)(α1+1)!(QsZ1s)α1+1|rds.

    Then, by the assumptions (A1),(A3),(A5), (A6) and Lemma 3.1, we obtain

    ttkE|X1(s,Qs,Z1s,Z2s)|rds 2r1ttk[E(|f(s,Qs,Z2s)|2)r2+Kr1[(a1+1)!]rE|QsZ1s|(α1+1)r]ds 2r1ttk[2r1KrE(|Qs|r+|Z2s|r)+Kr12(α1+1)r1[(a1+1)!]r(E|Qs|(α1+1)r+|Z1s|(α1+1)r)]ds 2r1ttk[2rKrK4+Kr12(α1+1)r[(a1+1)!]rK4]ds C2Δ. (3.6)

    In the same way as (3.6) was derived, we can show that

    ttkEX2(s,Qs,Z1s,,Z2s)r2dsC3Δ (3.7)

    and

    ttkE|X3(s,Qs,Z1s,Z2s)|rdsC4Δ. (3.8)

    Substituting (3.5), (3.6), (3.7) and (3.8) into (3.4) yields

    E|QtZt|r 5r1[K5Δr+K4(μ01ˉμ)rΔr+C2Δr+C3Δr2+C2C42r1Δr2+C42r1λrΔr] CΔr2,

    which is the required inequality (3.3).

    Lemma 3.3. Under the assumptions (A1)-(A6), Lemma 3.1 and E|Qsa|r<K5 holds. For 2r(α+1)q, there exists a γ(0,1] such that

    E|QtτZ2t|rCΔγ,t0. (3.9)

    Proof. For any t0, there exists an integer k0 such that t[tk,tk+1). We divide the whole proof into the following three cases.

    ● If τtkτtτ0. Then, by the assumption (A4), we have

    E|QtτZ2t|r=E|QtτQtkτ|r=E|ϕtτϕtkτ|r˜KΔγ. (3.10)

    ● If 0tkτtτ. Then, by Lemma 3.2, we have

    E|QtτZ2t|rCΔr2. (3.11)

    ● If τtkτ0tτ. Then, we have

    E|QtτZ2t|r2r1E|Qtτϕ0|r+2r1E|Qtkτϕ0|r. (3.12)

    Then together with (3.10) and (3.11), we have the following results immediately,

    E|QtτZ2t|rC(Δr2+Δγ). (3.13)

    Summarizing the above three cases, we therefore derive that

    E|QtτZ2t|rCΔγ

    for 2r(α+1)q and γ(0,1], which is the desired inequality (3.9).

    We can now begin to prove the following theorem which reveals the convergence of the Taylor approximate solutions to the true solutions.

    Theorem 3.1. , Let the assumptions (A1)(A6) and Lemma 3.1 hold. Then for any q2 and γ(0,1],

    Esup0tT|PtQt|qCΔγ. (3.14)

    Consequently

    limΔ0E[sup0tT|PtQt|q]=0. (3.15)

    Proof. By the (2.2) and (3.1), it is not difficult to show that

    PtQt=t0(PsQs)adst0(μe+(μ01μe)eηs)μ2(a)(PsQs)ds+ttk(f(s,Ps,Psτ)X1(s,Qs,Z1s,Z2s))ds+ttk(g(s,Ps,Psτ)X2(s,Qs,Zs,Z2s))dBs+ttk(h(s,Ps,Psτ)X3(s,Qs,Zs,Z2s))dNs.

    We write

    e(t)=PtQt,I1(s)=f(s,Ps,Psτ)X1(s,Qs,Z1s,Z2s),I2(s)=g(s,Ps,Psτ)X2(s,Qs,Z1s,Z2s),I3(s)=h(s,Ps,Psτ)X2(s,Qs,Z1s,Z2s)

    for simplicity. For all t[0,T], using Itô's formula to |e(t)|q and copying the analysis of (2.11) to (2.13), we have

    |e(t)|q (qˉβ2A2+2μ01ˉμq)2t0|e(s)|qds+t0q|e(s)|q1|I1(s)|ds+q(q1)2t0|e(s)|q2I2(s)22ds+t0q|e(s)|q1|I2(s)|dBs+t0q|e(s)|q1|I3(s)|dˉNs+q(q1)2t0|e(s)|q2|I3(s)|2dˉNs+λqt0|e(s)|q1|I3(s)|ds+λq(q1)2t0|e(s)|q2|I3(s)|2ds.

    The Young inequality yields

    |a|c|b|d|a|c+d+dc+d[cε(c+d)]cd|b|c+d (3.16)

    for a,bR and c,d,ε>0. We hence have

    Esup0tT|e(t)|q K6Esup0tTt0|e(s)|qds+Esup0tTt0q(|e(s)|q+K7|I1(s)|q)ds+q(q1)2Esup0tTt0(|e(s)|q+K8I2(s)q2)ds+λq(q1)2Esup0tTt0(|e(s)|q+K8|I3(s)|q)ds.+λqEsup0tTt0(|e(s)|q+K7|I3(s)|q)ds+q(q1)2Esup0tTt0|e(s)|q2|I3(s)|2dˉNs+Esup0tTt0q|e(s)|q2(e(s),I2(s))dBs+Esup0tTt0q|e(s)|q2(e(s),I3(s))dˉNs, (3.17)

    where K6=(qˉβ2A2+2μ01ˉμq)2, K7=1q[q1εq]q1 and K8=2q[q2εq]q22.

    Applying the Burkholder-Davis-Gundy's inequality [29] and Young inequality, we obtain that

    Esup0tTt0|e(s)|q2(e(s),I2(s))dBs 16qE[sup0tT|e(t)|q]+C1E(t0|e(s)|q2I2(s)22ds) 16qE[sup0tT|e(t)|q]+C1Esup0tTt0(|e(s)|q+K8I2(s)q2)ds. (3.18)

    Continuing this approach, we have

    Esup0tTt0|e(s)|q2(e(s),I3(s))dˉNs16qE[sup0tT|e(t)|q]+C2Esup0tTt0(|e(s)|q+K8|I3(s)|q)ds (3.19)

    and

    Esup0tTt0|e(s)|q2|I3(s)|2dˉNs13q(q1)E[sup0tT|e(t)|q]+C3Esup0tTt0(|e(s)|q+K9|I3(s)|q)ds, (3.20)

    where K9=4q[q4εq]q44.

    Next, under the assumptions (A3), (A6), Lemma 3.2 and Lemma 3.3, we then compute

    Esup0tTt0|I1(s)|qds= Esup0tTt0|f(s,Ps,Psτ)X1(s,Qs,Z1s,Z2s)|qds= Esup0tTt0|f(s,Ps,Psτ)f(s,Qs,Z2s)+f(s,Qs,Z2s)X1(s,Qs,Z1s,Z2s)|qds 2q1E[T0(|f(s,Ps,Psτ)f(s,Qs,Z2s)|q+|f(s,Qs,Z2s)X1(s,Qs,Z1s,Z2s)|q)ds] 2q1[22q2KqT0E(|PsQs|q+|PsτQsτ|q+|QsτZ2s|q)ds+T0E|f(α1+1)P(s,Z1s+θ(QsZ1s),Z2s)(a1+1)!(QsZ1s)α1+1|qds] 23q2Kq[T0Esup0us|e(u)|qds]+23q3KqTCΔγ+2q1Kq1TC[(a1+1)!]qΔ(α1+1)q2. (3.21)

    Moreover, we can similarly compute

    Esup0tTt0|I2(s)|qds= Esup0tTt0(g(s,Ps,Psτ)X2(s,Qs,Z1s,Z2s))q2ds 23q2Kq[T0Esup0us|e(u)|qds]+23q3KqTCΔγ+2q1Kq2TC[(a2+1)!]qΔ(α2+1)q2 (3.22)

    and

    Esup0tTt0|I3(s)|qds= Esup0tTt0|h(s,Ps,Psτ)X3(s,Qs,Z1s,Z2s)|qds 23q2Kq[T0Esup0us|e(u)|qds]+23q3KqTCΔγ+2q1Kq3TC[(a3+1)!]qΔ(α3+1)q2. (3.23)

    Substituting (3.18) to (3.23) into (3.17) we obtain that

    Esup0tT|e(t)|qCΔ(α+1)q2+CΔγ+Ct0Esup0us|e(u)|qds

    and the required result (3.14) and (3.15) follows from the Gronwall inequality.

    In this section, we present some numerical experiments to demonstrate the theoretical result. Let us consider the following age-dependent stochastic delay population equations with OU process and Poisson jumps:

    {dtPt=[Pta(0.65+(0.50.65)e0.75t)11aPt+f(t,Pt,Pt1)]dt+g(t,Pt,Pt1)dBt+h(t,Pt,Pt1)dNt,     (t,a)(0,1)×(0,1),P(t,a)=exp(11a), (t,a)[1,0]×[0,1],P(t,0)=A01(1a)2P(t,a)da,    t[0,1], (4.1)

    where A=1, T=1, τ=1, μe = 0.65, μ01 = 0.5, η = 0.75, μ2(a)=11a, Bt is a scalar Brownian motion, Nt is a Poisson process with intensity λ=1, ϕ(t,a)=exp(11a) and β(t,a)=1(1a)2.

    Now, we employ MATLAB for numerical simulations. First, we compare the convergence speed of the Taylor approximation scheme and backward Euler methods (BEM) mentioned in [30]. Let T=1, Δt=5×104 and Δa=0.05. For f, g and h of model (4.1), we choose three different groups of functions as examples. Obviously, it is easy to verify that assumptions (A1)(A6) are satisfied. By averaging over all of the 500 samples, on the computer running at Intel Core i5-4570 CPU 3.20 GHz, the runtimes of the Taylor approximation scheme (where f, g and h are approximated up to the 5th order) and the backward Euler methods for model (4.1) are given in Table 1.

    Table 1.  Runtimes for the Taylor approximation scheme and backward Euler method.
    f(t,Pt,Ptτ) g(t,Pt,Ptτ) h(t,Pt,Ptτ) Taylor approximation BEM
    sin2Pt+14sin4Pt1 P2t+1+Pt1 sinPtPt1 17.561282 29.227642
    expPt+P2t11 (lnPt)1P2t1 cosPt+sin2Pt1 19.325948 31.497162
    2PtlnPt+P2t1 sin3PtcosPt+Pt1 2PtlnPt+1 25.367845 34.894251

     | Show Table
    DownLoad: CSV

    Form the fist group f, g and h functions in Table 1, we observe that the runtime of the Taylor approximation scheme (3.1) is about 17.561282 seconds while the runtime of backward Euler method is about 29.227642 seconds on the same computer, and conclude that the convergence speed of the Taylor approximation scheme is 1.664 times faster than that of the backward Euler methods. As the theoretical results, Table 1 reveals that the rate of convergence for the Taylor approximation scheme is faster than the backward Euler methods.

    Next, we explore the convergence of the Taylor approximation scheme (3.1). By Theorem 3.1, we obtain that the numerical solution of the Taylor approximation scheme will converge to the exact solution with γq, where γ(0,1] and q2. Since the age-dependent stochastic delay population equations with OU process and Poisson jumps (4.1) cannot be solved analytically, we use more precise numerical solutions to obtain the exact solution. We take T=1, Δt=0.005, Δa=0.05, f(t,Pt,Ptτ)=sin2Pt+14sin4Pt1, g(t,Pt,Ptτ)=P2t+1+Pt1 and h(t,Pt,Ptτ)=sinPtPt1. Based on [5], the "explicit solutions" P(t,a) to model (4.1) can be given by the numerical solution of the SSθ method with θ=0.2.

    In Figure 1, we show the paths of the "explicit solutions" P(t,a), the numerical solution Q(t,a) of the Taylor approximation scheme (where f, g and h approximated up to the 10th order) and error simulations between them. In Figure 2, the relative difference between "explicit solutions" P(t, a) and numerical solutions Q(t,a) is presented. Moreover, we can see that the maximum value of the error square is less than 0.04 from Figure 1 and the maximum value of the the relative difference is less than 0.2 from Figure 2. Clearly the numerical solution Q(t,a) converge to exact solution in the mean sense.

    Figure 1.  The upper left corner shows the path of the "explicit solutions" P(t, a). The upper right corner displays the path of the Taylor approximation of solution of Q(t, a). The lower left and lower right diagrams represent mean and mean-square error simulations between "explicit solutions" P(t, a) and numerical solutions Q(t,a) based on the Taylor approximation, respectively.
    Figure 2.  The relative difference between "explicit solutions" P(t, a) and numerical solutions Q(t,a).

    To further demonstrate the convergence of the Taylor approximation scheme, we show the errors between the "explicit solutions" P(t,a) and numerical solutions Q(t,a) at different values of time step Δt, expansion order x and age step Δa in Table 2. Table 2(a) shows that for fixed expansion order x=10 and time step Δt=0.005, the corresponding value of (P(t,a)Q(t,a))2 when Δa take 0.04, 0.05, 0.2, 0.25 and 0.5, separately. In Table 2(b), for Taylor approximations of the coefficients f, g and h, we choose the expansion order to take 5, 8, 10, 15 and 20, separately. Then the values of (P(t,a)Q(t,a))2 are given. In Table 2(c), for fixed expansion order x=10 and age step Δa=0.05, we give the corresponding value of (P(t,a)Q(t,a))2 when Δt take 0.005, 0.001, 0.0005, 0.0001 and 0.00005, separately. To illustrate our results more succinctly and forcefully, we use log-log plot Figure 3(a)–(c) to simulate the data of Table 2(a)–(c), respectively. In Figure 3(b), when expansion order x=10 and time step Δt=0.005, as the value of age step Δa increases, the value of (P(t,a)Q(t,a))2 increases. In Figure 3(b), as one would expect, as the expansion order increases, the value of (P(t,a)Q(t,a))2 is getting smaller with time step Δt=0.005, age step Δa=0.05. From Figure 3(c), it clearly reveals the fact that for fixed expansion order x=10 and age step Δa=0.05, the value of (P(t,a)Q(t,a))2 will tend to decrease when the increments of time Δt smaller. Thus, based on the above numerical analysis, we conclude that the Taylor approximation scheme is a simple and efficient numerical method for the age-dependent stochastic delay population equations with OU process and Poisson jumps.

    Table 2.  Error simulation between P(t,a) and Q(t,a) at different values of Δa, x, and Δt.
    (a) x=10,Δt=0.005
    Δa0.04 0.05 0.2 0.25 0.5
    (P(t,a)Q(t,a))2 0.03 0.04 0.07 0.08 0.2
    (b)Δt=0.005,Δa=0.05
    x 5 8 10 20 50
    (P(t,a)Q(t,a))2 0.08 0.07 0.04 0.01 0.005
    (c)x=10,Δa=0.05
    Δt 0.005 0.001 0.0005 0.0001 0.00005
    (P(t,a)Q(t,a))2 0.04 0.04 0.03 0.02 0.01

     | Show Table
    DownLoad: CSV
    Figure 3.  Error simulation between P(t,a) and Q(t,a) at different values of order Δa, x, and Δt.

    This paper discuss a Taylor approximation scheme for a class of stochastic age-dependent population equations. In order to obtain a more realistic and improved model compared to those in the literature [1,2,3,4,5,6,7,8,9,11], we introduce the mean-reverting Ornstein-Uhlenbeck (OU) process, time delay and Poisson jumps into equation and form a new system (2.7). We investigate the pth moments boundedness of exact solutions of age-dependent stochastic delay population equations with mean-reverting OU process and Poisson jumps (2.7). When the drift and diffusion coefficients satisfies Taylor approximations, we construct a Taylor approximation scheme for Eq (2.7) and reveal that the Taylor approximation solutions converge to the exact solutions for the equations. Furthermore, we estimate the order of the convergence. We also utilize a numerical example to confirm our theoretical results. In our future work, we will consider the effect of variable delay for stochastic age-dependent population equations and investigate the convergence of numerical methods for stochastic age-dependent population equations with OU process and variable delay.

    The authors are very grateful to the anonymous reviewers for their insightful comments and helpful suggestions. This research was funded by the "Major Innovation Projects for Building First-class Universities in China's Western Region" (ZKZD2017009).

    All authors declare no conflicts of interest in this paper.

    Proof of Lemma 3.1. For the purpose of simplification, let l=(α+1)2q. Applying Itô's formula to |Qt|l, we have

    |Qt|l= |Q0|l+t0l|Qs|l2Qsa(μe+(μ01μe)eηs)μ2(a)Qs,Qsds+t0l|Qs|l2(α1j=0f(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j,Qs)ds+t0l(l1)2|Qs|l2α2j=0g(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j22ds+t0l|Qs|l2(Qs,α2j=0g(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)dBs+t0l|Qs|l2(Qs,α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)dNs+t0l(l1)2|Qs|l2α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j22dNs= |Q0|l+t0l|Qs|l2Qsa(μe+(μ01μe)eηs)μ2(a)Qs,Qsds+t0l|Qs|l2(α1j=0f(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j,Qs)ds+t0l(l1)2|Qs|l2α2j=0g(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j22ds+t0l|Qs|l2(Qs,α2j=0g(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)dBs+t0l|Qs|l2(Qs,α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)dˉNs+λt0l|Qs|l2(Qs,α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)ds+t0l(l1)2|Qs|l2|α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j|2dˉNs+λt0l(l1)2|Qs|l2|α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j|2ds, (5.1)

    where ˉNs=Nsλs is a compensated Poisson process. Since

    Qsa,Qs=A0Qsda(Qs)= 12(A0β(t,a)Qsda)2 12A0β2(t,a)daA0Q2sda 12ˉβ2A2|Qs|2, (5.2)

    by the assumptions (A1)-(A3), we get that

    |Qt|l |Q0|l+l(ˉβ2A22+μ01ˉμ)t0|Qs|lds+t0l|Qs|l2(α1j=0f(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j,Qs)ds+t0l(l1)2|Qs|l2α2j=0g(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j22ds+t0l|Qs|l2(Qs,α2j=0g(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)dBs+t0l|Qs|l2(Qs,α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)dˉNs+λt0l|Qs|l2(Qs,α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j)ds+t0l(l1)2|Qs|l2|α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j|2dˉNs+λt0l(l1)2|Qs|l2|α3j=0h(j)Z1s(s,Z1s,Z2s)j!(QsZ1s)j|2ds. (5.3)

    Using the well-known mean value theorem, we derive that there exists a θ(0,1) such that

    |Qt|l |Q0|l+l(ˉβ2A22+μ01ˉμ)t0|Qs|lds+t0l|Qs|l2(f(s,Qs,Z2s)f(α1+1)Z1s(s,Z1s+θ(QsZ1s),Z2s)(α1+1)!(QsZ1s)α1+1,Qs)ds+t0l(l1)2|Qs|l2g(s,Qs,Z2s)g(α2+1)Z1s(s,Z1s+θ(QsZ1s),Z2s)(α2+1)!(QsZ1s)α2+122ds
    +t0l|Qs|l2(Qs,g(s,Qs,Z2s)g(α2+1)Z1s(s,Z1s+θ(QsZ1s),Z2s)(α2+1)!(QsZ1s)α2+1)dBs+t0l|Qs|l2(Qs,h(s,Qs,Z2s)h(α3+1)Z1s(s,Z1s+θ(QsZ1s),Z2s)(α3+1)!(QsZ1s)α3+1)dˉNs+λt0l|Qs|l2(Qs,h(s,Qs,Z2s)h(α3+1)Z1s(s,Z1s+θ(QsZ1s),Z2s)(α3+1)!(QsZ1s)α3+1)ds+t0l(l1)2|Qs|l2|h(s,Qs,Z2s)h(α3+1)Z1s(s,Z1s+θ(QsZ1s),Z2s)(α3+1)!(QsZ1s)α3+1|2dˉNs+λt0l(l1)2|Qs|l2|h(s,Qs,Z2s)h(α3+1)Z1s(s,Z1s+θ(QsZ1s),Z2s)(α3+1)!(QsZ1s)α3+1|2ds. (5.4)

    Denotes

    H1(s,Qs,Z1s,Z2s)=f(α1+1)Z1s(s,Z1s+θ(QsZ1s),Z2s),H2(s,Qs,Z1s,Z2s)=g(α2+1)Z1s(s,Z1s+θ(QsZ1s),Z2s),H3(s,Qs,Z1s,Z2s)=h(α3+1)Z1s(s,Z1s+θ(QsZ1s),Z2s).

    Therefore, we obtain that

    t0l|Qs|l2(f(s,Qs,Z2s)H1(s,Qs,Z1s,Z2s)(α1+1)!(QsZ1s)α1+1,Qs)ds+t0l(l1)2|Qs|l2g(s,Qs,Z2s)H2(s,Qs,Z1s,Z2s)(α2+1)!(QsZ1s)α2+122ds+λt0l|Qs|l2(Qs,h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1)ds+λt0l(l1)2|Qs|l2|h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1|2ds t0l|Qs|l2(K(Qs+Z2s)+K1(α1+1)!2Qsα1+1,Qs)ds+t0l(l1)|Qs|l2(2K2Qs2+(K2(α2+1)!2α2+1)2Qs2α2+2)ds+λt0l(l1)|Qs|l2(2K2Qs2+(K3(α3+1)!2α3+1)2Qs2α3+2)ds+λt0l|Qs|l2(K(Qs+Z2s)+K3(α3+1)!2Qsα3+1,Qs)ds (1+λ)t0l|Qs|l2(2KQs2+K1(α1+1)!2α+1Qsα+2)ds+(1+λ)t0(2l(l1)K2|Qs|l+l(l1)(C(ˉα+1)!2α+1)2Qsl+2α)ds (1+λ)t0((2lK+2l(l1)K2)|Qs|l+(lK1(α1+1)!2α+1+l(l1)(C(ˉα+1)!2α+1)2)Qsl+2α)ds, (5.5)

    where ˉα=min{α1,α2,α3}.

    By the Burkholder-Davis-Gundy's inequality, we then have

    E[sup0stt0l|Qs|l2(Qs,g(s,Qs,Z2s)H2(s,Qs,Z1s,Z2s)(α2+1)!(QsZ1s)α2+1)dBs]+E[sup0stt0l|Qs|l2(Qs,h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1)dˉNs]+E[sup0stt0l(l1)2|Qs|l2|h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1|2dˉNs] E[sup0stt0l|Qs|l2(Ql22s,g(s,Qs,Z2s)H2(s,Qs,Z1s,Z2s)(α2+1)!(QsZ1s)α2+1)dBs]+E[sup0stt0l|Qs|l2(Ql22s,h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1)dˉNs]+E[sup0stt0l(l1)2|Qs|l2|h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1|2dˉNs] lE[supτut|Qu|l2(t0(Ql22s,g(s,Qs,Z2s)H2(s,Qs,Z1s,Z2s)(α2+1)!(QsZ1s)α2+1)dBs)]+lE[supτut|Qu|l2(t0(Ql22s,h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1)dˉNs)]+E[sup0stt0l(l1)2|Qs|l2|h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1|2dˉNs] 3lE[supτut|Qu|l2(t0|Qs|l2g(s,Qs,Z2s)H2(s,Qs,Z1s,Z2s)(α2+1)!(QsZ1s)α2+122ds)12]+3lE[supτut|Qu|l2(t0|Qs|l2h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+122ds)12]+E[sup0stt0l(l1)2|Qs|l2|h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1|2dˉNs] 3lE[supτut|Qu|l2(t0|Qs|l2(2K2Qs2+(K2(α2+1)!2α2+1)2Qs2α2+2)ds)12]+3lE[supτut|Qu|l2(t0|Qs|l2(2K2Qs2+(K3(α3+1)!2α3+1)2Qs2α3+2)ds)12]+E[sup0stt0l(l1)2|Qs|l2|h(s,Qs,Z2s)H3(s,Qs,Z1s,Z2s)(α3+1)!(QsZ1s)α3+1|2dˉNs] 16E[supτut|Qu|l]+CE(t0|Qs|l2(2K2Qs2+(K2(α2+1)!2α2+1)2Qs2α2+2)ds)+16E[supτut|Qu|l]+CE(t0|Qs|l2(2K2Qs2+(K3(α3+1)!2α3+1)2Qs2α3+2)ds)+16E[supτut|Qu|l]+CE(t0|Qs|l2(2K2Qs2+(K3(α3+1)!2α3+1)2Qs2α3+2)ds) 12E[supτut|Qu|l]+CE(t0(|Qs|l+Qs2α+l)ds). (5.6)

    Note that for any t[0,T],

    E[supτut|Qu|l]=E[supτu0|Qu|l]E[sup0ut|Qu|l].

    Combining (5.4), (5.5) and (5.6), we can show that

    E[supτut|Qu|l] E[supτu0|ϕu|l]+l(ˉβ2A22+μ01ˉμ+(1+λ)(2K+2(l1)K2))t0Esupτus|Qu|lds+(1+λ)(lK1(α1+1)!2α+1+l(l1)(C(ˉα+1)!2α+1)2)t0EsupτusQul+2αds+12E[supτut|Qu|l]+C(t0(Esupτus|Qu|l+EsupτusQu2α+l)ds) E[supτu0|ϕu|l]+Ct0Esupτus|Qu|lds+Ct0(EsupτusQu2α+l)ds E[supτu0|ϕu|l]+Ct0(EsupτusQu2l)ds.

    Finally, by Generalization of the Bellman lemma [31], we obtain the desired result (3.2).



    [1] M. K. Ali, The quantum normal form and its equivalents, J. Math. Phys., 26 (1985), 2565–2572. http://dx.doi.org/10.1063/1.526775 doi: 10.1063/1.526775
    [2] V. I. Arnold, Mathematical methods of classical mechanics, New York, NY: Springer, 1989. http://dx.doi.org/10.1007/978-1-4757-2063-1
    [3] D. V. Anosov, S. Kh. Aranson, V. I. Arnold, I. U. Bronshtein, V. Z. Grines, Yu. S. Ilyashenko, Dynamical systems I: ordinary differential equations and smooth dynamical systems, Berlin Heidelberg: Springer, 1988.
    [4] V. I. Arnold, Dynamical systems III: classical mechanics, Berlin, Heidelberg: Springer, 1988. http://dx.doi.org/10.1007/978-3-662-02535-2
    [5] V. I. Arnold, Dynamical systems V: bifurcation theory and catastrophe theory, Berlin, Heidelberg: Springer, 1994. http://dx.doi.org/10.1007/978-3-642-57884-7
    [6] D. Boccaletti, G. Pucacco, Theory of orbits, Berlin, Heidelberg: Springer, 1999. http://dx.doi.org/10.1007/978-3-662-09240-8
    [7] M. Born, P. Jordan, On quantum mechanics, Zeit. Phys., 34 (1925), 858–888.
    [8] M. Born, W. Heisenberg, P. Jordan, On quantum mechanics Ⅱ, Zeit. Phys., 35 (1926), 557–615.
    [9] S. Buyukdagli, M. Joyeux, On the application of canonical perturbation theory up to the dissociation threshold, Chem. Phys. Lett., 412 (2005), 200–205. http://dx.doi.org/10.1016/j.cplett.2005.06.115 doi: 10.1016/j.cplett.2005.06.115
    [10] H. W. Broer, G. A. Lunter, G. Vegter, Equivariant singularity theory with distinguished parameters: Two case studies of resonant Hamiltonian systems, Physica D, 112 (1998), 64–80. http://dx.doi.org/10.1016/S0167-2789(97)00202-9 doi: 10.1016/S0167-2789(97)00202-9
    [11] M. Cargo, A. Gracia-Saz, R. G. Littlejohn, M. W. Reinsch, P. de M Rios, Quantum normal forms, Moyal star product and Bohr-Sommerfeld approximation, J. Phys. A: Math. Gen., 38 (2005), 1977–2004. http://dx.doi.org/10.1088/0305-4470/38/9/010 doi: 10.1088/0305-4470/38/9/010
    [12] B. V. Chirikov, A universal instability of many-dimensional oscillator systems, Phys. Rep., 52 (1979), 263–379. http://dx.doi.org/10.1016/0370-1573(79)90023-1 doi: 10.1016/0370-1573(79)90023-1
    [13] G. Cicogna, G. Gaeta, Symmetry and perturbation theory in nonlinear dynamics, Berlin, Heidelberg: Springer. 1999. http://dx.doi.org/10.1007/3-540-48874-X
    [14] G. Contopoulos, Order and chaos in dynamical astronomy, Berlin, Heidelberg: Springer, 2002. http://dx.doi.org/10.1007/978-3-662-04917-4
    [15] R. H. Cushman, L. M. Bates, Global aspects of classical integrable systems, Basel: Birkhauser, 1997. http://dx.doi.org/10.1007/978-3-0348-8891-2
    [16] R. H. Cushman, H. R. Dullin, A. Giacobbe, D. D. Holm, M. Joyeux, P. Lynch, et al., CO2 molecule as a quantum realization of the 1:1:2 resonant swing-spring with monodromy, Phys. Rev. Lett., 93 (2004), 024302. http://dx.doi.org/10.1103/physrevlett.93.024302 doi: 10.1103/physrevlett.93.024302
    [17] A. S. Davydov, Quantum mechanics, Pergamon, 1965.
    [18] P. A. M. Dirac, The fundamental equations of quantum mechanics, Proc. R. Soc. Lond. A, 109 (1925), 642–653. http://dx.doi.org/10.1098/rspa.1925.0150 doi: 10.1098/rspa.1925.0150
    [19] B. Eckhardt, Birkhoff-Gustavson normal form in classical and quantum mechanics, J. Phys. A: Math. Gen., 19 (1986), 2961. http://dx.doi.org/10.1088/0305-4470/19/15/020 doi: 10.1088/0305-4470/19/15/020
    [20] K. Efstathiou, Metamorphoses of Hamiltonian systems with symmetries, Berlin, Heidelberg: Springer, 2005. http://dx.doi.org/10.1007/b105138
    [21] K. Efstathiou, D. A. Sadovskii, B. I. Zhilinskii, Classification of perturbations of the hydrogen atom by small static electric and magnetic fields, Proc. R. Soc. A, 463 (2007), 1771–1790. http://dx.doi.org/10.1098/rspa.2007.1843 doi: 10.1098/rspa.2007.1843
    [22] S. C. Farantos, R. Schinke, H. Guo, M. Joyeux, Energy localization in molecules, bifurcation phenomena, and their spectroscopic signatures: the global view, Chem. Rev., 109 (2009), 4248–4271. http://dx.doi.org/10.1021/cr900069m doi: 10.1021/cr900069m
    [23] F. Faure, B. I. Zhilinskii, Qualitative features of intra-molecular dynamics. What can be learned from symmetry and topology?, Acta Appl. Math., 70 (2002), 265–282. http://dx.doi.org/10.1023/A:1013986518747 doi: 10.1023/A:1013986518747
    [24] G. Gaeta, Lie-Poincaré transformations and a reduction criterion in Landau theory, Ann. Phys., 312 (2004), 511–540. http://dx.doi.org/10.1016/j.aop.2004.04.001 doi: 10.1016/j.aop.2004.04.001
    [25] G. Gaeta, Poincaré-like approach to Landau theory. Ⅰ. General theory, J. Math. Phys., 56 (2015), 083504. http://dx.doi.org/10.1063/1.4927425 doi: 10.1063/1.4927425
    [26] G. Gaeta, Poincaré-like approach to Landau theory. Ⅱ. Simplifying the Landau-deGennes potential for nematic liquid crystals, J. Math. Phys., 56 (2015), 083505. http://dx.doi.org/10.1063/1.4927426 doi: 10.1063/1.4927426
    [27] G. Gaeta, On the isotropic-biaxial phase transition in nematic liquid crystals, EPL-Europhys. Lett., 112 (2015), 46002. http://dx.doi.org/10.1209/0295-5075/112/46002 doi: 10.1209/0295-5075/112/46002
    [28] A. Giorgilli, Notes on Hamiltonian Dynamical Systems, Cambridge University Press, 2022.
    [29] M. Golubitsky, D. G. Schaeffer, Singularities and groups in bifurcation theory, New York, NY: Springer, 1985. http://dx.doi.org/10.1007/978-1-4612-5034-0
    [30] S. Graffi, Th. Paul, Convergence of a quantum normal form and an exact quantization formula, J. Funct. Anal., 262 (2012), 3340–3393. http://dx.doi.org/10.1016/j.jfa.2012.01.010 doi: 10.1016/j.jfa.2012.01.010
    [31] V. Guillemin, Th. Paul, Some remarks about semiclassical trace invariants and quantum normal forms, Commun. Math. Phys., 294 (2010), 1. http://dx.doi.org/10.1007/s00220-009-0920-3 doi: 10.1007/s00220-009-0920-3
    [32] H. Hanßmann, Local and semi-local bifurcations in Hamiltonian dynamical systems, Berlin, Heidelberg: Springer, 2007. http://dx.doi.org/10.1007/3-540-38894-X
    [33] M. Hénon, C. Heiles, The applicability of the third integral of motion: some numerical experiments, Astron. J. 69 (1964), 73–79.
    [34] M. Herman, D. S. Perry, Molecular spectroscopy and dynamics: a polyad-based perspective, Phys. Chem. Chem. Phys., 15 (2013), 9970–9993. http://dx.doi.org/10.1039/C3CP50463H doi: 10.1039/C3CP50463H
    [35] J. Henrard, Periodic orbits emanating from a resonant equilibrium, Celestial Mechanics, 1 (1970), 437–466. http://dx.doi.org/10.1007/BF01231143 doi: 10.1007/BF01231143
    [36] D. Hilbert, Theory of algebraic invariants, Cambridge: Cambridge University Press, 1993.
    [37] T. Iwai, B. I. Zhilinskii, Local description of band rearrangements, Acta Appl. Math., 137 (2015), 97–121. http://dx.doi.org/10.1007/s10440-014-9992-y doi: 10.1007/s10440-014-9992-y
    [38] M. P. Jacobson, R. W. Field, Acetylene at the threshold of isomerization, J. Phys. Chem. A, 104 (2000), 3073–3086. http://dx.doi.org/10.1021/jp992428u doi: 10.1021/jp992428u
    [39] M. Joyeux, Gustavson's procedure and the dynamics of highly excited vibrational states, J. Chem. Phys., 109 (1998), 2111–2122. http://dx.doi.org/10.1063/1.476724 doi: 10.1063/1.476724
    [40] M. Joyeux, D. Sugny, M. Lombardi, Vibrational dynamics up to the dissociation threshold: A case study of two-dimensional HOCI, J. Chem. Phys., 113 (2000), 9610–9621. http://dx.doi.org/10.1063/1.1321031 doi: 10.1063/1.1321031
    [41] M. Joyeux, D. Sugny, Canonical perturbation theory for highly excited dynamics, Can. J. Phys., 80 (2002), 1459–1480. http://dx.doi.org/10.1139/p02-075 doi: 10.1139/p02-075
    [42] S. V. Krasnoshchekov, N. F. Stepanov, Polyad quantum numbers and multiple resonances in anharmonic vibrational studies of polyatomic molecules, J. Chem. Phys., 139 (2013), 184101. http://dx.doi.org/10.1063/1.4829143 doi: 10.1063/1.4829143
    [43] M. Kummer, On resonant non linearly coupled oscillators with two equal frequencies, Commun. Math. Phys., 48 (1976), 53–79. http://dx.doi.org/10.1007/BF01609411 doi: 10.1007/BF01609411
    [44] L. D. Landau, E. M. Lifshitz, Quantum mechanics, Pergamon Press, 1958.
    [45] A. Marchesiello, G. Pucacco, Equivariant singularity analysis of the 2:2 resonance, Nonlinearity, 27 (2014), 43–66. http://dx.doi.org/10.1088/0951-7715/27/1/43 doi: 10.1088/0951-7715/27/1/43
    [46] A. B. McCoy, E. L. Sibert, Perturbative calculations of vibrational (J = 0) energy levels of linear molecules in normal coordinate representations, J. Chem. Phys., 95 (1991), 3476–3487. http://dx.doi.org/10.1063/1.460850 doi: 10.1063/1.460850
    [47] A. B. McCoy, D. C. Burleigh, E. L. Sibert, Rotation-vibration interactions in highly excited states of SO2 and H2CO, J. Chem. Phys., 95 (1991), 7449–7465. http://dx.doi.org/10.1063/1.461371 doi: 10.1063/1.461371
    [48] A. B. McCoy, E. L. Sibert, The bending dynamics of acetylene, J. Chem. Phys., 105 (1996), 459–468. http://dx.doi.org/10.1063/1.471899 doi: 10.1063/1.471899
    [49] A. Messiah, Quantum mechanics, Wiley, 1958.
    [50] K. R. Meyer, G. R. Hall, D. C. Offin, Introduction to Hamiltonian dynamical systems and the N-body problem, Berlin: Springer, 2009.
    [51] J. Montaldi, M. Roberts, I. Stewart, Existence of nonlinear normal modes of symmetric Hamiltonian systems, Nonlinearity, 3 (1990), 695–730. http://dx.doi.org/10.1088/0951-7715/3/3/009 doi: 10.1088/0951-7715/3/3/009
    [52] J. Montaldi, M. Roberts, I. Stewart, Stability of nonlinear normal modes of symmetric Hamiltonian systems, Nonlinearity, 3 (1990), 731–772. http://dx.doi.org/10.1088/0951-7715/3/3/010 doi: 10.1088/0951-7715/3/3/010
    [53] J. Moser, Regularization of Kepler's problem and the averaging method on a manifold, Commun. Pure Appl. Math., 23 (1970), 609–636. http://dx.doi.org/10.1002/cpa.3160230406 doi: 10.1002/cpa.3160230406
    [54] A. I. Neishtadt, Averaging and passage through resonance, In: Proceedings of the International Congress of Mathematicians, Kyoto, Japan, 1990, 1271–1283.
    [55] A. I. Neishtadt, Averaging, passage through resonances, and capture into resonance in two-frequency systems, Russ. Math. Surv., 69 (2014), 771. http://dx.doi.org/10.1070/RM2014v069n05ABEH004917 doi: 10.1070/RM2014v069n05ABEH004917
    [56] S. V. Ngoc, Quantum Birkhoff normal forms and semiclassical analysis, Adv. Stud. Pure Math., 55 (2009), 99–116. http://dx.doi.org/10.2969/aspm/05510099 doi: 10.2969/aspm/05510099
    [57] I. M. Pavlichenkov, Bifurcations in quantum rotational spectra, Phys. Rep., 226 (1993), 173–279. http://dx.doi.org/10.1016/0370-1573(93)90083-P doi: 10.1016/0370-1573(93)90083-P
    [58] I. M. Pavlichenkov, B. I. Zhilinskii, Critical phenomena in rotational spectra, Ann. Phys., 184 (1988), 1–32. http://dx.doi.org/10.1016/0003-4916(88)90268-0 doi: 10.1016/0003-4916(88)90268-0
    [59] L. E. Picasso, Lectures in quantum mechanics, Cham: Springer, 2016. http://dx.doi.org/10.1007/978-3-319-22632-3
    [60] G. Pierre, D. A. Sadovskii, B. I. Zhilinskii, Organization of quantum bifurcations: crossover of rovibrational bands in spherical top molecules, EPL-Europhys. Lett., 10 (1989), 409–414. http://dx.doi.org/10.1209/0295-5075/10/5/004 doi: 10.1209/0295-5075/10/5/004
    [61] H. Poincaré, Les méthodes nouvelles de la méchanique céleste, Paris: Gauthier-Villars, 1892.
    [62] G. Pucacco, A. Marchesiello, An energy-momentum map for the time-reversal symmetric 1:1 resonance with Z2×Z2 symmetry, Physica D, 271 (2014), 10–18. http://dx.doi.org/10.1016/j.physd.2013.12.009 doi: 10.1016/j.physd.2013.12.009
    [63] J. P. Rose, M. E. Kellman, Bending dynamics from acetylene spectra: Normal, local, and precessional modes, J. Chem. Phys., 105 (1996), 10743–10754. http://dx.doi.org/10.1063/1.472882 doi: 10.1063/1.472882
    [64] S. Sachdev, Quantum phase transitions, Cambridge: Cambridge University Press, 1999. http://dx.doi.org/10.1017/CBO9780511622540
    [65] D. A. Sadovskii, B. I. Zhilinskii, Counting levels within vibrational polyads: Generating function approach, J. Chem. Phys., 103 (1995), 10520. http://dx.doi.org/10.1063/1.469836 doi: 10.1063/1.469836
    [66] D. A. Sadovskii, B. I. Zhilinskii, Quantum monodromy, its generalizations and molecular manifestations, Mol. Phys., 104 (2006), 2595–2615. http://dx.doi.org/10.1080/00268970600673363 doi: 10.1080/00268970600673363
    [67] D. A. Sadovskii, B. I. Zhilinskii, Hamiltonian systems with detuned 1:1:2 resonance: Manifestations of bidromy, Ann. Phys., 322 (2007), 164–200. http://dx.doi.org/10.1016/j.aop.2006.09.011 doi: 10.1016/j.aop.2006.09.011
    [68] J. A. Sanders, F. Verhulst, J. Murdock, Averaging methods in nonlinear dynamical systems, Berlin: Springer, 2007. http://dx.doi.org/10.1007/978-0-387-48918-6
    [69] M. Sanrey, M. Joyeux, Quantum mechanical and quasiclassical investigations of the time domain nonadiabatic dynamics of NO2 close to the bottom of the X2A1A2B2 conical intersection, J. Chem. Phys., 125 (2006), 014304. http://dx.doi.org/10.1063/1.2211609 doi: 10.1063/1.2211609
    [70] M. Sanrey, M. Joyeux, Slow periodic oscillations in time domain dynamics of NO2, J. Chem. Phys., 126 (2007), 074301. http://dx.doi.org/10.1063/1.2446920 doi: 10.1063/1.2446920
    [71] D. S. Schmidt, Periodic solutions near a resonant equilibrium of a Hamiltonian system, Celestial Mechanics, 9 (1974), 81–103. http://dx.doi.org/10.1007/BF01236166 doi: 10.1007/BF01236166
    [72] S. Schmidt, H. R. Dullin, Dynamics near the p:q resonance, Physica D, 239 (2010), 1884–1891. http://dx.doi.org/10.1016/j.physd.2010.06.012 doi: 10.1016/j.physd.2010.06.012
    [73] I. Shavitt, L. T. Redmon, Quasidegenerate perturbation theories. A canonical van Vleck formalism and its relationship to other approaches, J. Chem. Phys., 73 (1980), 5711–5717. http://dx.doi.org/10.1063/1.440050 doi: 10.1063/1.440050
    [74] E. L. Sibert, Theoretical studies of vibrationally excited polyatomic molecules using canonical Van Vleck perturbation theory, J. Chem. Phys., 88 (1988), 4378–4390. http://dx.doi.org/10.1063/1.453797 doi: 10.1063/1.453797
    [75] E. L. Sibert, A. B. McCoy, Quantum, semiclassical and classical dynamics of the bending modes of acetylene, J. Chem. Phys., 105 (1996), 469–478. http://dx.doi.org/10.1063/1.471900 doi: 10.1063/1.471900
    [76] J. F. Svitak, V. Tyng, M. E. Kellman, Bifurcation analysis of higher m:n resonance spectroscopic Hamiltonian, J. Phys. Chem. A, 106 (2002), 10797–10805. http://dx.doi.org/10.1021/jp0263976 doi: 10.1021/jp0263976
    [77] J. M. Tuwankotta, F. Verhulst, Symmetry and resonance in Hamiltonian systems, SIAM J. Appl. Math., 61 (2000), 1369–1385. http://dx.doi.org/10.1137/S0036139900365323 doi: 10.1137/S0036139900365323
    [78] F. Verhulst, Discrete symmetric dynamical systems at the main resonances with applications to axi-symmetric galaxies, Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences, 290 (1979), 435–465. http://dx.doi.org/10.1098/rsta.1979.0006 doi: 10.1098/rsta.1979.0006
    [79] A. Weinstein, Normal modes for nonlinear Hamiltonian systems, Invent. Math., 20 (1973), 47–57. http://dx.doi.org/10.1007/BF01405263 doi: 10.1007/BF01405263
    [80] B. I. Zhilinskii, Symmetry, invariants, and topology in molecular models, Phys. Rep., 341 (2001), 85–171. http://dx.doi.org/10.1016/S0370-1573(00)00089-2 doi: 10.1016/S0370-1573(00)00089-2
    [81] B. I. Zhilinskii, Rearrangement of energy bands: topological aspects, J. Math. Chem., 44 (2008), 1009–1022. http://dx.doi.org/10.1007/s10910-008-9359-6 doi: 10.1007/s10910-008-9359-6
    [82] B. I. Zhilinskii, Quantum bifurcations, New York: Springer, 2009. http://dx.doi.org/10.1007/978-0-387-30440-3
    [83] B. I. Zhilinskii, G. Dhont, D. A. Sadovskii, New Qualitative Phenomena in Classical (Intramolecular) Dynamics and their Quantum Analogs, 2007. Available from: https://kb.osu.edu/handle/1811/31349.
  • This article has been cited by:

    1. Huili Wei, Wenhe Li, Dynamical behaviors of a Lotka-Volterra competition system with the Ornstein-Uhlenbeck process, 2023, 20, 1551-0018, 7882, 10.3934/mbe.2023341
    2. Meng Gao, Xiaohui Ai, A stochastic Gilpin-Ayala nonautonomous competition model driven by mean-reverting OU process with finite Markov chain and Lévy jumps, 2024, 32, 2688-1594, 1873, 10.3934/era.2024086
    3. Meng Gao, Xiaohui Ai, A stochastic Gilpin-Ayala mutualism model driven by mean-reverting OU process with Lévy jumps, 2024, 21, 1551-0018, 4117, 10.3934/mbe.2024182
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2394) PDF downloads(218) Cited by(0)

Figures and Tables

Figures(9)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog