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

Lyapunov Exponents for Hamiltonian Systems under Small Lévy Perturbations

Ying Chao1, Pingyuan Wei2,∗ and Jinqiao Duan3 1 School of Mathematics and Statistics, Xi’an Jiaotong University, Xi’an 710049, China 2School of Mathematics and Statistics & Center for Mathematical Sciences, Huazhong University of Science and Technology, Wuhan 430074, China 3Department of Applied Mathematics, Illinois Institute of Technology, Chicago, IL 60616, USA Corresponding author: [email protected] [email protected], [email protected], [email protected].
Abstract

This work is to investigate the (top) Lyapunov exponent for a class of Hamiltonian systems under small non-Gaussian Lévy noise. In a suitable moving frame, the linearisation of such a system can be regarded as a small perturbation of a nilpotent linear system. The Lyapunov exponent is then estimated by taking a Pinsky-Wihstutz transformation and applying the Khas’minskii formula, under appropriate assumptions on smoothness, ergodicity and integrability. Finally, two examples are present to illustrate our results. The result characterizes the growth or decay rates of a class of dynamical systems under the interaction between Hamiltonian structures and non-Gaussian uncertainties.

Keywords: Lyapunov exponent; Hamiltonian systems; (non-Gaussian) Lévy noise; Pinsky-Wihstutz transformation; stochastic stability along a trajectory.

1 Introduction

The randomly influenced Hamiltonian systems are mathematical models for complex phenomena in physical, chemical and biological science [11, 6, 13]. These systems are stochastic differential equations (SDEs) whose deterministic counterparts have Hamiltonian structures, and their solutions are referred as (Hamiltonian) diffusion processes. As random fluctuation (also called, random noise) is present, the original deterministic Hamiltonian structure is usually modified or destroyed, and some interesting phenomena may occur. A meaningful issue is to investigate the stochastic stability of these systems by evaluating the Lyapunov exponents of the corresponding linearized systems.

For stochastic Hamiltonian systems driven by (Gaussian) Brownian motion, this stability problem has been studied extensively, for example, in Arnold [6], Arnold-Oeljeklaus-Pardoux [7], Ariaratnam-Xie [5], Baxendale-Goukasian [9, 10] and Zhu [25]. There are two common ways to linearize these systems. One way is to consider the linearizations at equilibrium states and get the corresponding (top) Lyapunov exponents. Then the signs of Lyapunov exponents determine the (local) stabilities for the original systems. Within a certain region of an equilibrium state, consider two different trajectories obtained by starting the original SDE at distinct points. If the Lyapunov exponent is negative, it is clear that they will simultaneously converge to the equilibrium state. But, if the Lyapunov exponent is positive, it is not readily apparent what will happen to the distance apart of these two trajectories as time goes to infinity. This gives rise to the other way to linearize systems (i.e., along trajectories), and then examine the stochastic stability along trajectories. In this present paper, we will adopt the latter one. For more related results on Lyapunov exponents of SDEs with Brownian motion, we refer to [8, 16].

However, the Gaussian paradigm is known to be too limited in many physical contexts. Non-Gaussian random fluctuations, including heavy-tailed distributions and burst-like events as in earthquakes or abrupt climate change, are widely observed in scientific and engineering systems [21, 4, 12]. Recently, Lyapunov exponents of stochastic differential equations with (non-Gaussian) Lévy noise have received increasing attention, as in, for example, Mao-Rodkina [17], Applebaum [4] and Qiao-Duan [20]. It is thus desirable to investigate Lyapunov exponents for Hamiltonian system evolutions under non-Gaussian Lévy fluctuations. But very few works are available on this topic as far as we know.

Our goal in this paper is to estimate the (top) Lyapunov exponents for one degree-of-freedom Hamiltonian systems under small Lévy perturbations. Based on interlacing technique [4], the Lévy perturbations consider here consist of a Brownian part and a bounded jump part. These systems should be interpreted as modified versions of Marcus SDEs [18, 15] and the solution processes are jump diffusion processes. One of our main inspirations comes from the work [10] by Baxendale and Goukasian, where the authors studied a class of stochastic Duffing-van der Pol equations, and estimated the Lyapunov exponents for this Hamiltonian system under small Gaussian perturbation. We generalize and improve their approach to deal with the case of non-Gaussian Lévy perturbations. This gives rise to several difficulties both in analytic and probabilistic aspects.

This paper is organized as follows. In Section 2, we recall some basic facts on Lévy motions and Hamiltonian systems with Lévy noise. Section 3 is dedicated to formulate precise assumptions and present our main results, Theorem 3.3. We first show that the linearized equations are randomly influenced nilpotent linear systems. Then we introduce a Pinsky-Wihstutz transformation and apply a Khas’minskii formula to estimate the Lyapunov exponent. Finally, we present two specific illustrative examples in Section 4.

2 Preliminaries

We first recall some basic facts on Lévy motion [4, 21, 12], and introduce a class of stochastic Hamiltonian systems to be studied.

2.1 Lévy Motions

Let (Ω,,{t}t0,)(\Omega,\mathscr{F},\{\mathscr{F}_{t}\}_{t\geqslant 0},\mathbb{P}) be a filtered probability space, where t\mathscr{F}_{t} is a nondecreasing family of sub-σ\sigma-fields of \mathscr{F} satisfying the usual conditions. An t\mathscr{F}_{t}-adapted stochastic process Lt=L(t)L_{t}=L(t) taking values in d\mathbb{R}^{d} with L(0)=0L(0)=0 a.s.a.s. (almost surely) is called a Lévy motion if it is stochastically continuous, with independent increments and stationary increments.

An dd-dimensional Lévy motion can be characterized by a drift vector bdb\in\mathbb{R}^{d}, an d×dd\times d non-negative-definite, symmetric matrix QQ, and a Borel measure ν\nu defined on d\{0}{\mathbb{R}^{d}}\backslash\{0\}. We call (b,Q,ν)(b,Q,\nu) the generating triplet of the Lévy motion LtL_{t} and write this Lévy motion as Lt(b,Q,ν)L_{t}\sim(b,Q,\nu) for convenient. Moreover, we have the following Lévy-Itô decomposition for LtL_{t}:

Lt=bt+BQ(t)+z<czN~(t,dz)+zczN(t,dz),{L_{t}}=bt+B_{Q}(t)+\int_{\|z\|<c}z\tilde{N}(t,dz)+\int_{\|z\|\geq c}zN(t,dz), (2.1)

where N(dt,dz)N(dt,dz) is the Poisson random measure on +×(d\{0})\mathbb{R}^{+}\times({\mathbb{R}^{d}}\backslash\{0\}), N~(dt,dz)=N(dt,dz)ν(dz)dt\tilde{N}(dt,dz)=N(dt,dz)-\nu(dz)dt is the compensated Poisson random measure, ν=𝔼N(1,)\nu=\mathbb{E}N(1,\cdot) is the jump measure, BQ(t)B_{Q}(t) is an independent dd-dimensional Brownian motion with covariance matrix QQ, and the last two terms describe the ‘small jumps’ and ‘big jumps’ of Lévy process, respectively. Here \|\cdot\| denotes the Euclidean norm and cc is a positive constant. In the following, we denote Lc(t)=γt+BQ(t)L_{c}(t)=\gamma t+B_{Q}(t) as the continuous part of LtL_{t} and Ld(t)=LtLc(t)L_{d}(t)=L_{t}-L_{c}(t) as the discontinuous part.

2.2 Hamiltonian Systems with Lévy noise

Let H:2H:\mathbb{R}^{2}\to\mathbb{R} be a smooth function with isolated critical points such that H(x)H(x)\to\infty as x\|x\|\to\infty, and U1=(H/x2,H/x1)TU_{1}=(\partial H/\partial x^{2},-\partial H/\partial x^{1})^{T} be the Hamiltonian vector field with respect to HH. Given smooth vector fields V1,,VdV_{1},\cdots,V_{d} and a dd-dimensional Lévy motion Lt=(Ltk)k=1,,dL_{t}=(L^{k}_{t})_{k=1,\cdots,d} with generating triplet (0,I,ν)(0,I,\nu). For 0<ε<10<\varepsilon<1, we consider the following one degree-of-freedom Hamiltonian system with Lévy noise

dxt=U1(xt)dt+εk=1dVk(xt)dLtk,x(0)=x0,dx_{t}=U_{1}(x_{t})dt+\varepsilon\sum_{k=1}^{d}V_{k}(x_{t})\diamond dL^{k}_{t},\;\;x(0)=x_{0}, (2.2)

or equivalently,

xt=x0+0tU1(xs)𝑑s+εk=1d0tVk(xs)dLsk,x_{t}=x_{0}+\int_{0}^{t}U_{1}(x_{s})ds+\varepsilon\sum_{k=1}^{d}\int_{0}^{t}V_{k}(x_{s-})\diamond dL^{k}_{s}, (2.3)

where “\diamond” indicates Marcus (canonical) integral [18, 15] defined by

εk=1d0tVk(xs)dLsk=\displaystyle\varepsilon\sum_{k=1}^{d}\int_{0}^{t}V_{k}(x_{s-})\diamond dL^{k}_{s}= εk=1d0tVk(xs)𝑑Lc,sk+εk=1d0tVk(xs)𝑑Ld,sk\displaystyle\varepsilon\sum_{k=1}^{d}\int_{0}^{t}V_{k}(x_{s-})\circ dL^{k}_{c,s}+\varepsilon\sum_{k=1}^{d}\int_{0}^{t}V_{k}(x_{s-})dL^{k}_{d,s}
+0st[ξε(ΔLs(xs))xsεk=1dVk(xs)ΔLsk]\displaystyle+\sum_{0\leqslant s\leqslant t}\left[\xi^{\varepsilon}(\Delta L_{s}(x_{s-}))-x_{s-}-\varepsilon\sum_{k=1}^{d}V_{k}(x_{s-})\Delta L^{k}_{s}\right]

with 𝑑Lc,sk\int\cdots\circ dL^{k}_{c,s} denoting the Stratonovtich integral, 𝑑Ld,sk\int\cdots dL^{k}_{d,s} denoting the Itô integral and ξε\xi^{\varepsilon} being the value at τ=1\tau=1 of the solution of the following ODE: (for each x2x\in\mathbb{R}^{2} and zdz\in\mathbb{R}^{d})

dξε(τz)(x)dτ=εk=1dzkVk(ξε(τz)(x)),ξε(0)(x)=x.\displaystyle\frac{d\xi^{\varepsilon}(\tau z)(x)}{d\tau}=\varepsilon\sum_{k=1}^{d}z_{k}V_{k}(\xi^{\varepsilon}(\tau z)(x)),\;\;\xi^{\varepsilon}(0)(x)=x. (2.4)

Compared with Itô stochastic differentials, Marcus one has the advantage of leading to ordinary chain rule under a transformation (change of variable). This offers some slight reduction in the lengths of some of the calculations, but has no serious effect on the theorem. We also remark that (2.2) is referred as a stochastic Hamiltonian system preserving symplectic structure if VkV_{k}, k=1,dk=1,\cdots d are Hamiltonian vector fields [23].

Rewritting SDE (2.2) into Itô form, we have

dxt=\displaystyle dx_{t}= U1(xt)dt+12ε2k=1d(DVkVk)(xt)dt+εk=1dVk(xt)dBtk\displaystyle U_{1}(x_{t})dt+\frac{1}{2}\varepsilon^{2}\sum_{k=1}^{d}(DV_{k}V_{k})(x_{t})dt+\varepsilon\sum_{k=1}^{d}V_{k}(x_{t})dB_{t}^{k}
+z<c[ξε(z)(xt)xtεk=1dzkVk(xt)]ν(dz)𝑑t\displaystyle+\int_{\|z\|<c}\left[\xi^{\varepsilon}(z)(x_{t-})-x_{t-}-\varepsilon\sum_{k=1}^{d}z_{k}V_{k}(x_{t-})\right]\nu(dz)dt
+z<c[ξε(z)(xt)xt]N~(dt,dz)\displaystyle+\int_{\|z\|<c}\left[\xi^{\varepsilon}(z)(x_{t-})-x_{t-}\right]\tilde{N}(dt,dz)
+zc[ξε(z)(xt)xt]N(dt,dz),\displaystyle+\int_{\|z\|\geqslant c}\left[\xi^{\varepsilon}(z)(x_{t-})-x_{t-}\right]{N}(dt,dz), (2.5)

where cc is a positive constant. The term in (2.2) involving large jump is controlled by a function on {zc}{\{\|z\|\geqslant c\}}, and it would be absent if we take c=c=\infty. One standard way to handle this case is to use interlacing technique [4]. In this paper, we use this technique directly without making precise statements. For more details about interlacing, we would like to refer to Applebaum [4]. Hence, we can omit the large-jump terms and concentrate on the study of the equations driven by continuous noise interspersed with small jumps, then SDE (2.2) can be modified as

dxt=\displaystyle dx_{t}= U1(xt)dt+12ε2k=1d(DVkVk)(xt)dt+εk=1dVk(xt)dBtk\displaystyle U_{1}(x_{t})dt+\frac{1}{2}\varepsilon^{2}\sum_{k=1}^{d}(DV_{k}V_{k})(x_{t})dt+\varepsilon\sum_{k=1}^{d}V_{k}(x_{t})dB_{t}^{k}
+z<c[ξε(z)(xt)xtεk=1dzkVk(xt)]ν(dz)𝑑t\displaystyle+\int_{\|z\|<c}\left[\xi^{\varepsilon}(z)(x_{t-})-x_{t-}-\varepsilon\sum_{k=1}^{d}z_{k}V_{k}(x_{t-})\right]\nu(dz)dt
+z<c[ξε(z)(xt)xt]N~(dt,dz).\displaystyle+\int_{\|z\|<c}\left[\xi^{\varepsilon}(z)(x_{t-})-x_{t-}\right]\tilde{N}(dt,dz). (2.6)

If we assume that U1(x)U_{1}(x) and V~(x)k=1d(DVkVk)(x)\tilde{V}(x)\triangleq\sum_{k=1}^{d}(DV_{k}V_{k})(x) are locally Lipschitz continuous functions and satisfy ‘one sided linear growth’ condition in the following sense:

  1. (A1)

    (Locally Lipschitz condition) For any R>0R>0, there exists K1>0K_{1}>0 such that, for all x1\|x_{1}\|, x2R\|x_{2}\|\leqslant R,

    U1(x1)U1(x2)2+V~(x1)V~(x2)2+max1kdVk(x1)Vk(x2)2K1x1x22,\|U_{1}(x_{1})-U_{1}(x_{2})\|^{2}+\|\tilde{V}(x_{1})-\tilde{V}(x_{2})\|^{2}+\max_{1\leqslant k\leqslant d}\|V_{k}(x_{1})-V_{k}(x_{2})\|^{2}\leqslant K_{1}\|x_{1}-x_{2}\|^{2},
  2. (A2)

    (One sided linear growth condition) There exists K2>0K_{2}>0 such that, for all x2x\in\mathbb{R}^{2},

    k=1dVk(x)2+2x(U1+V~)(x)K2(1+x2),\sum_{k=1}^{d}\|V_{k}(x)\|^{2}+2x\cdot(U_{1}+\tilde{V})(x)\leqslant K_{2}(1+\|x\|^{2}),

then there exists a unique solution to (2.2) and the solution process is adapted and cádlàg, based on Lemma 6.10.3 of [4] and Theorem 3.1 of [1]. Referring to Theorem 6.8.2 of [4], assumption (A2) also ensures that the solution process of (2.2) has a Lyapunov exponent.

3 Lyapunov Exponent for A Hamiltonian System under Small Lévy Perturbation

From now on, we focus on the modified SDE (2.2), which is indeed a one degree-of-freedom Hamiltonian system with small Lévy perturbation. For convenient, we rewrite SDE (2.2) into canonical form:

dxt=U1(xt)dt+εk=1dVk(xt)dL~tkdx_{t}=U_{1}(x_{t})dt+\varepsilon\sum_{k=1}^{d}V_{k}(x_{t})\diamond d\tilde{L}^{k}_{t} (3.1)

with L~tk\tilde{L}^{k}_{t} a Lévy-type noise neglecting large jump part. Linearizing the canonical SDE (3.1) along the trajectory xtx_{t}, we get the following equation

dvt=DU1(xt)vtdt+εk=1dDVk(xt)vtdL~tk.dv_{t}=DU_{1}(x_{t})v_{t}dt+\varepsilon\sum_{k=1}^{d}DV_{k}(x_{t})v_{t}\diamond d\tilde{L}^{k}_{t}. (3.2)

Instead of calculate the Lyapunov exponent of such a linearization system directly, we would like to estimate it by taking advantage of the specific structure of equation (3.2).

3.1 Nilpotent Structure of the Linearization System

We first claim that, when rewritten with respect to a suitable moving frame, the linearization of perturbed Hamiltonian system (3.1), that is, system (3.2), is a perturbed nilpotent linear system. For convenience, we write U2(x)=H(x)/H(x)2,U_{2}(x)=\nabla H(x)/\|\nabla H(x)\|^{2}, and adopt the notation U.f(x)=Df(x)(U(x))=f(x),U(x)U.f(x)=Df(x)(U(x))=\langle\nabla f(x),U(x)\rangle, which is the action of a vector field UU as a first order differential operator acting on a function ff.

Denote by MM the space 2\mathbb{R}^{2} with all critical points of HH removed. Note that, for xMx\in M, (U1.H)(x)=0(U_{1}.H)(x)=0, (U2.H)(x)=1(U_{2}.H)(x)=1 and U1(x),U2(x)=0\langle U_{1}(x),U_{2}(x)\rangle=0. We rewrite smooth vector fields VkV_{k}, k=1,,dk=1,\cdots,d, into

Vk(x)=ak1(x)U1(x)+ak2(x)U2(x),xM,{V}_{k}(x)=a_{k}^{1}(x)U_{1}(x)+a_{k}^{2}(x)U_{2}(x),\;\;x\in M, (3.3)

where ak1a_{k}^{1} and ak2a_{k}^{2} are smooth functions with respect to xx. Then the original system (3.1) becomes

dxt=U1(xt)dt+εk=1d(ak1U1+ak2U2)(xt)dL~tk.dx_{t}=U_{1}(x_{t})dt+\varepsilon\sum_{k=1}^{d}(a_{k}^{1}U_{1}+a_{k}^{2}U_{2})(x_{t})\diamond d\tilde{L}^{k}_{t}. (3.4)

Furthermore, we have the following lemma:

Lemma 3.1.

Let wt=(wt1,wt2)w_{t}=(w_{t}^{1},w_{t}^{2}) represent the linearized process vtv_{t} in a moving frame given by U1(x)U_{1}(x) and U2(x)U_{2}(x). That is,

vt=wt1U1(xt)+wt2U2(xt){v}_{t}=w_{t}^{1}U_{1}(x_{t})+w_{t}^{2}U_{2}(x_{t}) (3.5)

Define exist time τε=inft0{xtM or xt=±}\tau_{\varepsilon}=\inf_{t\geqslant 0}\{x_{t}\notin M\text{ or }x_{t}=\pm\infty\} as the first time that the process xtx_{t} given by (3.1) hits a critical point of HH or explodes to infinity. For all t<τεt<\tau_{\varepsilon}, we have

dwt=Λ(xt)wtdt+εk=1dMk(xt)wtdL~tk,dw_{t}=\Lambda(x_{t})w_{t}dt+\varepsilon\sum_{k=1}^{d}M_{k}(x_{t})w_{t}\diamond d\tilde{L}^{k}_{t}, (3.6)

where Λ=[0A00]\Lambda=\begin{bmatrix}\begin{matrix}0&A\\ 0&0\end{matrix}\end{bmatrix}, Mk=[BkCkDkEk]M_{k}=\begin{bmatrix}\begin{matrix}B_{k}&C_{k}\\ D_{k}&E_{k}\end{matrix}\end{bmatrix} (k=1,,d)(k=1,\cdots,d) with

A(x)={[((2H)2(1H)2)(22H11H)+41H2H12H]/H2}(x),\displaystyle A(x)=\big{\{}\big{[}\big{(}(\partial_{2}H)^{2}-(\partial_{1}H)^{2}\big{)}(\partial_{22}H-\partial_{11}H)+4\partial_{1}H\partial_{2}H\partial_{12}H\big{]}/\|\nabla H\|^{2}\big{\}}(x),
Bk(x)=(U1.ak1)(x)A(x)ak2(x),Dk(x)=(U1.ak2)(x),\displaystyle B_{k}(x)=(U_{1}.a_{k}^{1})(x)-A(x)a_{k}^{2}(x),\;\;\;\;\;\;\;D_{k}(x)=(U_{1}.a_{k}^{2})(x),
Ck(x)=(U2.ak1)(x)+A(x)ak1(x),Ek(x)=(U2.ak2)(x).\displaystyle C_{k}(x)={(U_{2}.a_{k}^{1})(x)+A(x)a_{k}^{1}(x)},\;\;\;\;\;\;\;E_{k}(x)=(U_{2}.a_{k}^{2})(x).
Proof.

The idea is to apply change of variable formula in the sense of canonical differential which could be referred to Proposition 4.3 of [15], and to follow the lines of [10, Lemma 3]. By substituting (3.3) in the equation (3.2), we have

dvt=DU1(xt)vtdt+εk=1di=12(akiDUi+Ui[aki]T)(xt)vtdL~tk.dv_{t}=DU_{1}(x_{t})v_{t}dt+\varepsilon\sum_{k=1}^{d}\sum_{i=1}^{2}\big{(}a_{k}^{i}DU_{i}+U_{i}[\nabla a_{k}^{i}]^{T}\big{)}(x_{t})v_{t}\diamond d\tilde{L}^{k}_{t}.

And then using (3.5), we obtain

dvt=\displaystyle dv_{t}= j=12wtjDU1Uj(xt)dt+εk=1di,j=12wtj(akiDUiUj+Uj.akiUi)(xt)dL~tk.\displaystyle\sum_{j=1}^{2}w_{t}^{j}DU_{1}U_{j}(x_{t})dt+\varepsilon\sum_{k=1}^{d}\sum_{i,j=1}^{2}w_{t}^{j}\big{(}a_{k}^{i}DU_{i}U_{j}+U_{j}.a_{k}^{i}U_{i}\big{)}(x_{t})\diamond d\tilde{L}^{k}_{t}. (3.7)

On the other hand, by differentiating (3.5) with respect to tt and taking (3.4) into account, we conclude that

dvt=\displaystyle d{v}_{t}= j=12dwtjUj(xt)+j=12wtjDUjdxt\displaystyle\sum_{j=1}^{2}dw_{t}^{j}U_{j}(x_{t})+\sum_{j=1}^{2}w_{t}^{j}DU_{j}dx_{t}
=\displaystyle= j=12dwtjUj(xt)+j=12wtjDUj(U1(xt)dt+εk=1di=12akiUi(xt)dL~tk)\displaystyle\sum_{j=1}^{2}dw_{t}^{j}U_{j}(x_{t})+\sum_{j=1}^{2}w_{t}^{j}DU_{j}\left(U_{1}(x_{t})dt+\varepsilon\sum_{k=1}^{d}\sum_{i=1}^{2}a_{k}^{i}U_{i}(x_{t})\diamond d\tilde{L}^{k}_{t}\right)
=\displaystyle= j=12dwtjUj(xt)+j=12wtj(DUjU1)(xt)dt+εk=1di,j=12wtj(akiDUjUi)(xt)dL~tk).\displaystyle\sum_{j=1}^{2}dw_{t}^{j}U_{j}(x_{t})+\sum_{j=1}^{2}w_{t}^{j}(DU_{j}U_{1})(x_{t})dt+\varepsilon\sum_{k=1}^{d}\sum_{i,j=1}^{2}w_{t}^{j}(a_{k}^{i}DU_{j}U_{i})(x_{t})\diamond d\tilde{L}^{k}_{t}). (3.8)

Note that (DU1U2DU2U1)(x)=(AU1)(x)(DU_{1}U_{2}-DU_{2}U_{1})(x)=(AU_{1})(x), as shown in Lemma 1 of [10]. We equate these two semimartingale expressions for dvtdv_{t}, that is, (3.7)-(3.1). The result follows by comparing their coefficient terms and directly calculation. ∎

3.2 A Lévy-type Pinsky-Wihstutz Transformation

Lemma 3.1 shows that we are indeed dealing with a small Lévy-type perturbation of a nilpotent system. Motivated by the remarkable work on small random perturbation of a nilpotent system in Pinsky-Wihstutz [19] and its applications in [9, 10], we define

T=[εβ001],  0<β<1,T=\begin{bmatrix}\begin{matrix}\varepsilon^{\beta}&0\\ 0&1\end{matrix}\end{bmatrix},\;\;0<\beta<1, (3.9)

for fixed ε>0\varepsilon>0. Since

min(εβ,1)wTwmax(εβ,1)w,\min(\varepsilon^{\beta},1)\|w\|\leqslant\|Tw\|\leqslant\max(\varepsilon^{\beta},1)\|w\|,

it follows that

limt1tlogTwt=limt1tlogwt\lim_{t\to\infty}\frac{1}{t}\log\|Tw_{t}\|=\lim_{t\to\infty}\frac{1}{t}\log\|w_{t}\|

and so the process TwtTw_{t} has the same Lyapunov exponent as wtw_{t}. For simplicity of notation we continue to write wtw_{t} in place of TwtTw_{t}.

Remark 3.2.

For Brownian case, that is, Lt(0,I,0)L_{t}\sim(0,I,0), the value of β\beta in transformation (3.9) is equal to 2/32/3 [19]. Such a transformation can be used to convert a problem involving singular perturbation into one involving a non-singular perturbation; see, for example, [9, 10]. Moreover, after the transformation, we have a pair process {(xt,wt):t0}\{(x_{t},w_{t}):t\geqslant 0\} with xtx_{t} moves at the fast rate 1 and wtw_{t} moving at slow rate εβ\varepsilon^{\beta}. It means that, assuming enough ergodicity and integrability, a stochastic averaging argument is possible.

Keeping in mind the equation for xtx_{t} in (2.2) or equivalently (3.1), the process wtw_{t} (t<τε)(t<\tau_{\varepsilon}) is now given by

dwt=εβΛ(xt)wtdt+εk=1dMkε(xt)wtdL~tk,dw_{t}=\varepsilon^{\beta}\Lambda(x_{t})w_{t}dt+\varepsilon\sum_{k=1}^{d}M_{k}^{\varepsilon}(x_{t})w_{t}\diamond d\tilde{L}^{k}_{t}, (3.10)

where Mkε=TMkT1=[BkεβCkεβDkEk]M_{k}^{\varepsilon}=TM_{k}T^{-1}=\begin{bmatrix}\begin{matrix}B_{k}&\varepsilon^{\beta}C_{k}\\ \varepsilon^{-\beta}D_{k}&E_{k}\end{matrix}\end{bmatrix} for k=1,,dk=1,\cdots,d.

Note that canonical differential satisfies the change of variable formula [15], we write

wt=wt[cosθtsinθt],w_{t}=\|w_{t}\|\begin{bmatrix}\begin{matrix}\cos\theta_{t}\\ \sin\theta_{t}\end{matrix}\end{bmatrix},

and define ρt=logwt\rho_{t}=\log\|w_{t}\|. Hence, equation (3.10) can be rewritten as a pair of equations:

dθt=\displaystyle d\theta_{t}= εβA(xt)sin2θt+k=1dσk1(xt,θt)dL~tk,\displaystyle-\varepsilon^{\beta}A(x_{t})\sin^{2}\theta_{t}+\sum_{k=1}^{d}\sigma_{k}^{1}(x_{t},\theta_{t})\diamond d\tilde{L}^{k}_{t}, (3.11)
dρt=\displaystyle d\rho_{t}= εβA(xt)sinθtcosθt+k=1dσk2(xt,θt)dL~tk,\displaystyle\varepsilon^{\beta}A(x_{t})\sin\theta_{t}\cos\theta_{t}+\sum_{k=1}^{d}\sigma_{k}^{2}(x_{t},\theta_{t})\diamond d\tilde{L}^{k}_{t}, (3.12)

where, for k=1,2,,dk=1,2,\cdots,d,

σk1(x,θ)=\displaystyle\sigma_{k}^{1}(x,\theta)= ε1βDk(x)cos2θε(BkEk)(x)sinθcosθε1+βCk(x)sin2θ\displaystyle\varepsilon^{1-\beta}D_{k}(x)\cos^{2}\theta-\varepsilon(B_{k}-E_{k})(x)\sin\theta\cos\theta-\varepsilon^{1+\beta}C_{k}(x)\sin^{2}\theta
\displaystyle\triangleq ε1βQ1k(x,θ)+εQ2k(x,θ)+ε1+βQ3k(x,θ),\displaystyle\varepsilon^{1-\beta}Q_{1}^{k}(x,\theta)+\varepsilon Q_{2}^{k}(x,\theta)+\varepsilon^{1+\beta}Q_{3}^{k}(x,\theta), (3.13)
σk2(x,θ)=\displaystyle\sigma_{k}^{2}(x,\theta)= ε1βDk(x)sinθcosθ+ε[Bk(x)cos2θ+Ek(x)sin2θ]+ε1+βCk(x)sinθcosθ\displaystyle\varepsilon^{1-\beta}D_{k}(x)\sin\theta\cos\theta+\varepsilon[B_{k}(x)\cos^{2}\theta+E_{k}(x)\sin^{2}\theta]+\varepsilon^{1+\beta}C_{k}(x)\sin\theta\cos\theta
\displaystyle\triangleq ε1βP1k(x,θ)+εP2k(x,θ)+ε1+βP3k(x,θ).\displaystyle\varepsilon^{1-\beta}P_{1}^{k}(x,\theta)+\varepsilon P_{2}^{k}(x,\theta)+\varepsilon^{1+\beta}P_{3}^{k}(x,\theta). (3.14)

Notice that, under the change of variables, the vector fields σki\sigma_{k}^{i}, k=1,,dk=1,\cdots,d, i=1,2i=1,2, should be understood as functions of xx, θ\theta and ρ\rho, but they indeed do not depend on ρ\rho. The fact that the equation for logwt\log\|w_{t}\| does not involve wt\|w_{t}\| was first used by Khas’minskii [14] in the 1960s, and is an essential part of the derivation of the Furstenberg-Khas’minskii formula for the top Lyapunov exponent of a linearized system.

We next further convert Marcus equations (3.11)-(3.12) into the Itô form [4, Page 418]. It is important to note that the Marcus interpretation for SDE (3.11)-(3.12) depends also on the SDE (3.4) for xtx_{t}. More specifically, the Marcus interpretation is only for SDE describing a diffusion process [15], and the pair process {(θt,ρt):t0}\{(\theta_{t},\rho_{t}):t\geqslant 0\} is not a diffusion process because its distribution depends also on {xt:t0}\{x_{t}:t\geqslant 0\}. A jump in the Lévy process LtαL_{t}^{\alpha} causes jumps in both xtx_{t} and (θt,ρt)(\theta_{t},\rho_{t}), we thus need to supplement ODE (2.4) with the following equations:

dζiε(τz)(x,θ)dτ=k=1dzkσki(ξε(τz)(x),ζ1ε(τz)(x,θ)),i=1,2,\displaystyle\frac{d\zeta_{i}^{\varepsilon}(\tau z)(x,\theta)}{d\tau}=\sum_{k=1}^{d}z_{k}\sigma_{k}^{i}\big{(}\xi^{\varepsilon}(\tau z)(x),\zeta_{1}^{\varepsilon}(\tau z)(x,\theta)\big{)},\;\;i=1,2, (3.15)

with ζε(0)(x,θ)=(x,θ)\zeta^{\varepsilon}(0)(x,\theta)=(x,\theta). That is, we need to focus on the time one flow ηε(z)(x,θ,ρ)=(ξε,ζ1ε,ρ+ζ2ε)\eta^{\varepsilon}(z)(x,\theta,\rho)=(\xi^{\varepsilon},\zeta_{1}^{\varepsilon},\rho+\zeta_{2}^{\varepsilon}) started at (x,θ,ρ)(x,\theta,\rho) along the vector field k=1dzkV~k(x,θ,ρ)\sum_{k=1}^{d}z_{k}\tilde{V}_{k}(x,\theta,\rho), where

V~k(x,θ,ρ)=(εVk(x),σk1(x,θ),σk2(x,θ)).\displaystyle\tilde{V}_{k}(x,\theta,\rho)=\Big{(}\varepsilon V_{k}(x),\sigma_{k}^{1}(x,\theta),\sigma_{k}^{2}(x,\theta)\Big{)}. (3.16)

Therefore, we have the following Itô SDE:

dθt=\displaystyle d\theta_{t}= εβA(xt)sin2θtdt+12k=1dσ~k1(xt,θt)dt+k=1dσk1(xt,θt)dBtk\displaystyle-\varepsilon^{\beta}A(x_{t})\sin^{2}\theta_{t}dt+\frac{1}{2}\sum_{k=1}^{d}\tilde{\sigma}_{k}^{1}(x_{t},\theta_{t})dt+\sum_{k=1}^{d}\sigma_{k}^{1}(x_{t},\theta_{t})dB_{t}^{k}
+z<c[ζ1ε(z)(xt,θt)θtk=1dzkσk1(xt,θt)]ν(dz)𝑑t\displaystyle+\int_{\|z\|<c}\left[\zeta_{1}^{\varepsilon}(z)(x_{t-},\theta_{t-})-\theta_{t-}-\sum_{k=1}^{d}z_{k}\sigma_{k}^{1}(x_{t-},\theta_{t-})\right]\nu(dz)dt
+z<c[ζ1ε(z)(xt,θt)θt]N~(dt,dz),\displaystyle+\int_{\|z\|<c}\left[\zeta_{1}^{\varepsilon}(z)(x_{t-},\theta_{t-})-\theta_{t-}\right]\tilde{N}(dt,dz), (3.17)
dρt=\displaystyle d\rho_{t}= εβA(xt)sinθtcosθtdt+12k=1dσ~k2(xt,θt)dt+k=1dσk2(xt,θt)dBtk\displaystyle\varepsilon^{\beta}A(x_{t})\sin\theta_{t}\cos\theta_{t}dt+\frac{1}{2}\sum_{k=1}^{d}\tilde{\sigma}_{k}^{2}(x_{t},\theta_{t})dt+\sum_{k=1}^{d}\sigma_{k}^{2}(x_{t},\theta_{t})dB_{t}^{k}
+z<c[ζ2ε(z)(xt,θt)k=1dzkσk2(xt,θt)]ν(dz)𝑑t\displaystyle+\int_{\|z\|<c}\left[\zeta_{2}^{\varepsilon}(z)(x_{t-},\theta_{t-})-\sum_{k=1}^{d}z_{k}\sigma_{k}^{2}(x_{t-},\theta_{t-})\right]\nu(dz)dt
+z<c[ζ2ε(z)(xt,θt)]N~(dt,dz)\displaystyle+\int_{\|z\|<c}\left[\zeta_{2}^{\varepsilon}(z)(x_{t-},\theta_{t-})\right]\tilde{N}(dt,dz) (3.18)

with Wong-Zakai correction terms

σ~k1(x,θ)=\displaystyle\tilde{\sigma}_{k}^{1}(x,\theta)= εDxσk1Vk+Dθσk1σk1\displaystyle\varepsilon D_{x}\sigma_{k}^{1}V_{k}+D_{\theta}\sigma_{k}^{1}\sigma_{k}^{1}
=\displaystyle= ε22βDθQ1kQ1k+ε2β(DxQ1kVk+DθQ1kP2k+DθQ2kQ1k)\displaystyle\varepsilon^{2-2\beta}D_{\theta}Q_{1}^{k}Q_{1}^{k}+\varepsilon^{2-\beta}(D_{x}Q_{1}^{k}V_{k}+D_{\theta}Q_{1}^{k}P_{2}^{k}+D_{\theta}Q_{2}^{k}Q_{1}^{k})
+ε2(DxQ2kVk+DθQ1kQ3k+DθQ3kQ1k)\displaystyle+\varepsilon^{2}(D_{x}Q_{2}^{k}V_{k}+D_{\theta}Q_{1}^{k}Q_{3}^{k}+D_{\theta}Q_{3}^{k}Q_{1}^{k})
+ε2+β(DxQ3kVk+DθQ2kQ3k+DθQ3kQ2k)+ε2+2βDθQ3kQ3k\displaystyle+\varepsilon^{2+\beta}(D_{x}Q_{3}^{k}V_{k}+D_{\theta}Q_{2}^{k}Q_{3}^{k}+D_{\theta}Q_{3}^{k}Q_{2}^{k})+\varepsilon^{2+2\beta}D_{\theta}Q_{3}^{k}Q_{3}^{k}
\displaystyle\triangleq ε22βQ~1k(x,θ)+ε2βQ~2k(x,θ)+ε2Q~3k(x,θ)+ε2+βQ~4k(x,θ)+ε2+2βQ~5k(x,θ),\displaystyle\varepsilon^{2-2\beta}\tilde{Q}_{1}^{k}(x,\theta)+\varepsilon^{2-\beta}\tilde{Q}_{2}^{k}(x,\theta)+\varepsilon^{2}\tilde{Q}_{3}^{k}(x,\theta)+\varepsilon^{2+\beta}\tilde{Q}_{4}^{k}(x,\theta)+\varepsilon^{2+2\beta}\tilde{Q}_{5}^{k}(x,\theta),
σ~k2(x,θ)=\displaystyle\tilde{\sigma}_{k}^{2}(x,\theta)= εDxσk2Vk+Dθσk2σk1\displaystyle\varepsilon D_{x}\sigma_{k}^{2}V_{k}+D_{\theta}\sigma_{k}^{2}\sigma_{k}^{1}
=\displaystyle= ε22βDθP1kQ1k+ε2β(DxP1kVk+DθP1kQ2k+DθP2kQ1k)\displaystyle\varepsilon^{2-2\beta}D_{\theta}P_{1}^{k}Q_{1}^{k}+\varepsilon^{2-\beta}(D_{x}P_{1}^{k}V_{k}+D_{\theta}P_{1}^{k}Q_{2}^{k}+D_{\theta}P_{2}^{k}Q_{1}^{k})
+ε2(DxP2kVk+DθP1kQ3k+DθP3kQ1k)\displaystyle+\varepsilon^{2}(D_{x}P_{2}^{k}V_{k}+D_{\theta}P_{1}^{k}Q_{3}^{k}+D_{\theta}P_{3}^{k}Q_{1}^{k})
+ε2+β(DxP5kVk+DθP2kQ3k+DθP3kQ2k)+ε2+2βDθP3kQ3k\displaystyle+\varepsilon^{2+\beta}(D_{x}P_{5}^{k}V_{k}+D_{\theta}P_{2}^{k}Q_{3}^{k}+D_{\theta}P_{3}^{k}Q_{2}^{k})+\varepsilon^{2+2\beta}D_{\theta}P_{3}^{k}Q_{3}^{k}
\displaystyle\triangleq ε22βP~1k(x,θ)+ε2βP~2k(x,θ)+ε2P~3k(x,θ)+ε2+βP~4k(x,θ)+ε2+2βP~5k(x,θ).\displaystyle\varepsilon^{2-2\beta}\tilde{P}_{1}^{k}(x,\theta)+\varepsilon^{2-\beta}\tilde{P}_{2}^{k}(x,\theta)+\varepsilon^{2}\tilde{P}_{3}^{k}(x,\theta)+\varepsilon^{2+\beta}\tilde{P}_{4}^{k}(x,\theta)+\varepsilon^{2+2\beta}\tilde{P}_{5}^{k}(x,\theta).

3.3 Estimating the Lyapunov Exponent

Based on Pinsky-Wihstutz transformation (3.9), we now proceed to estimate the Lyapunov exponent for the linearized version of the perturbed Hamiltonian system (3.1). We first note that, for the process {(xt,θt):t0}\{(x_{t},\theta_{t}):t\geqslant 0\} given by (2.2) and (3.17), its generator ε\mathcal{L}_{\varepsilon} can be written as

(εf)(x,θ)=\displaystyle(\mathcal{L}_{\varepsilon}f)(x,\theta)= [U1(x)+12ε2k=1d(DVkVk)(x)]xf(x,θ)\displaystyle\bigg{[}U_{1}(x)+\frac{1}{2}\varepsilon^{2}\sum_{k=1}^{d}(DV_{k}V_{k})(x)\bigg{]}\cdot\nabla_{x}f(x,\theta)
+[εβA(x)sin2θ+12k=1dσ~k1(x,θ)]θf(x,θ)\displaystyle+\bigg{[}-\varepsilon^{\beta}A(x)\sin^{2}\theta+\frac{1}{2}\sum_{k=1}^{d}\tilde{\sigma}_{k}^{1}(x,\theta)\bigg{]}\frac{\partial}{\partial\theta}f(x,\theta)
+12k=1d(ε2Tr[VkTVkxx]+2εVkσk1(x,θ)xθ+(σk1(x,θ))222θ)f(x,θ)\displaystyle+\frac{1}{2}\sum_{k=1}^{d}\bigg{(}\varepsilon^{2}\text{Tr}[V_{k}^{T}V_{k}\nabla_{x}\nabla_{x}]+2\varepsilon V_{k}\sigma_{k}^{1}(x,\theta)\cdot\nabla_{x}\frac{\partial}{\partial\theta}+({\sigma}_{k}^{1}(x,\theta))^{2}\frac{\partial^{2}}{\partial^{2}\theta}\bigg{)}f(x,\theta)
+z<1{f(ξε(z)(x),ζ1ε(z)(x,θ))f(x,θ)\displaystyle+\int_{\|z\|<1}\Bigg{\{}f\big{(}\xi^{\varepsilon}(z)(x),\zeta_{1}^{\varepsilon}(z)(x,\theta)\big{)}-f(x,\theta)
k=1dzk[εVk(x)xf(x,θ)+σk1(x,θ)θf(x,θ)]}ν(dz),\displaystyle\;\;\;-\sum_{k=1}^{d}z_{k}\Big{[}\varepsilon V_{k}(x)\cdot\nabla_{x}f(x,\theta)+\sigma_{k}^{1}(x,\theta)\frac{\partial}{\partial\theta}f(x,\theta)\Big{]}\Bigg{\}}\nu(dz), (3.19)

for each fCb2(M×S1)f\in C_{b}^{2}(M\times S^{1}) (i.e., ff is a C2C^{2} and bounded function). Referring to [3, 1] and keeping assumptions (A1)-(A2) in mind, we assume throughout the paper the following:

  1. (A3)

    For each sufficiently small ε>0\varepsilon>0, the process {(xt,θt):t0}\{(x_{t},\theta_{t}):t\geqslant 0\} given by (2.2) and (3.17) is a positive recurrent diffusion process on M×S1M\times S^{1} with a (unique) stationary probability με\mu^{\varepsilon}. We write μMϵ\mu_{M}^{\epsilon} for the MM marginal of με\mu^{\varepsilon}.

From the equation (3.18) for ρt\rho_{t}, we observe that the leading drift term (with respect to ε\varepsilon) of this equation, formally, not only depends on the value of β\beta, but also may contain contributions from the ν\nu-integral term. More specifically, it may come from the following terms: εβA(x)sinθcosθ\varepsilon^{\beta}A(x)\sin\theta\cos\theta, ε22β2k=1dP~1k(x,θ)\frac{\varepsilon^{2-2\beta}}{2}\sum_{k=1}^{d}\tilde{P}_{1}^{k}(x,\theta) and Iρεz<c[ζ2ε(z)(x,θ)k=1dzkσk2(x,θ)]ν(dz)I_{\rho}^{\varepsilon}\triangleq\int_{\|z\|<c}\left[\zeta_{2}^{\varepsilon}(z)(x,\theta)-\sum_{k=1}^{d}z_{k}\sigma_{k}^{2}(x,\theta)\right]\nu(dz).

As mentioned in Remark 3.2, for pure Brownian case, it has been verified that β=2/3\beta=2/3, so the leading drift term in pure Brownian case should be

ε2/3(A(x)sinθcosθ+1/2k=1dP~1k(x,θ))\displaystyle\varepsilon^{2/3}\Big{(}A(x)\sin\theta\cos\theta+1/2\sum_{k=1}^{d}\tilde{P}_{1}^{k}(x,\theta)\Big{)}
=ε2/3(A(x)sinθcosθ+k=1dDk2(x)(12cos2θsin2θcos2θ)).\displaystyle=\varepsilon^{2/3}\Big{(}A(x)\sin\theta\cos\theta+\sum_{k=1}^{d}D_{k}^{2}(x)(\frac{1}{2}\cos^{2}\theta-\sin^{2}\theta\cos^{2}\theta)\Big{)}. (3.20)

We next try to analyze the leading term of IρϵI_{\rho}^{\epsilon} formally. By equations (2.4) and (3.15), and following calculations in the proof for Lemma 6.10.3 of [4], we conclude that

Iρε(x,θ)=\displaystyle I_{\rho}^{\varepsilon}(x,\theta)= z<c[01k=1dzkσk2(ξε(τz)(x),ζ1ε(τz)(x,θ))dτk=1dzkσk2(x,θ)]ν(dz)\displaystyle\int_{\|z\|<c}\left[\int_{0}^{1}\sum_{k=1}^{d}z_{k}\sigma_{k}^{2}\big{(}\xi^{\varepsilon}(\tau z)(x),\zeta_{1}^{\varepsilon}(\tau z)(x,\theta)\big{)}d\tau-\sum_{k=1}^{d}z_{k}\sigma_{k}^{2}(x,\theta)\right]\nu(dz)
=\displaystyle= z<c010ak,l=1dzkzl(εDxσk2Vl+Dθσk2σl1)(ξε(bz)(x),ζ1ε(bz)(x,θ))dbdaν(dz).\displaystyle\int_{\|z\|<c}\int_{0}^{1}\int_{0}^{a}\sum_{k,l=1}^{d}z_{k}z_{l}\big{(}\varepsilon D_{x}\sigma_{k}^{2}V_{l}+D_{\theta}\sigma_{k}^{2}\sigma_{l}^{1}\big{)}\big{(}\xi^{\varepsilon}(bz)(x),\zeta_{1}^{\varepsilon}(bz)(x,\theta)\big{)}dbda\;\nu(dz).

Note that the expression for (εDxσk2Vl+Dθσk2σl1)(x,θ)\big{(}\varepsilon D_{x}\sigma_{k}^{2}V_{l}+D_{\theta}\sigma_{k}^{2}\sigma_{l}^{1}\big{)}(x,\theta) is quite similar to that of σ~k2(x,θ)\tilde{\sigma}_{k}^{2}(x,\theta). We adopt the following notation

Iρε(x,θ)z<c[ε22βR0ε+ε2βR1ε+ε2R2ε+ε2+βR3ε+ε2+2βR4ε](z)(x,θ)ν(dz),\displaystyle I_{\rho}^{\varepsilon}(x,\theta)\triangleq\int_{\|z\|<c}\left[\varepsilon^{2-2\beta}R_{0}^{\varepsilon}+\varepsilon^{2-\beta}R_{1}^{\varepsilon}+\varepsilon^{2}R_{2}^{\varepsilon}+\varepsilon^{2+\beta}R_{3}^{\varepsilon}+\varepsilon^{2+2\beta}R_{4}^{\varepsilon}\right](z)(x,\theta)\nu(dz), (3.21)

where, with Int[f]010ak,l=1dzkzlf(ξε(bz),ζ1ε(bz))dbdaInt[f]\triangleq\int_{0}^{1}\int_{0}^{a}\sum_{k,l=1}^{d}z_{k}z_{l}f(\xi^{\varepsilon}(bz),\zeta_{1}^{\varepsilon}(bz))dbda,

R0ε(z)\displaystyle R_{0}^{\varepsilon}(z) =Int[DθP1kQ1l],\displaystyle=Int[D_{\theta}P_{1}^{k}Q_{1}^{l}],\;\;\;
R1ε(z)\displaystyle R_{1}^{\varepsilon}(z) =Int[DxP1kVl+DθP1kQ2l+DθP2kQ1l]\displaystyle=Int[D_{x}P_{1}^{k}V_{l}+D_{\theta}P_{1}^{k}Q_{2}^{l}+D_{\theta}P_{2}^{k}Q_{1}^{l}]
R2ε(z)\displaystyle R_{2}^{\varepsilon}(z) =Int[DxP2kVl+DθP1kQ3l+DθP3kQ1l],\displaystyle=Int[D_{x}P_{2}^{k}V_{l}+D_{\theta}P_{1}^{k}Q_{3}^{l}+D_{\theta}P_{3}^{k}Q_{1}^{l}],\;\;\;
R3ε(z)\displaystyle R_{3}^{\varepsilon}(z) =Int[DxP5kVl+DθP2kQ3l+DθP3kQ2l]\displaystyle=Int[D_{x}P_{5}^{k}V_{l}+D_{\theta}P_{2}^{k}Q_{3}^{l}+D_{\theta}P_{3}^{k}Q_{2}^{l}]
R4ε(z)\displaystyle R_{4}^{\varepsilon}(z) =Int[DθP3kQ3l].\displaystyle=Int[D_{\theta}P_{3}^{k}Q_{3}^{l}].

In particular,

R0ε(z)=010a(k=1dzkDk(ξε(bz)))2[12cos2(ζ1ε(bz))sin2(ζ1ε(bz))cos2(ζ1ε(bz))]𝑑b𝑑a.\displaystyle R_{0}^{\varepsilon}(z)=\int_{0}^{1}\int_{0}^{a}\left(\sum_{k=1}^{d}z_{k}D_{k}\big{(}\xi^{\varepsilon}(bz)\big{)}\right)^{2}\left[\frac{1}{2}\cos^{2}\big{(}\zeta_{1}^{\varepsilon}(bz)\big{)}-\sin^{2}\big{(}\zeta_{1}^{\varepsilon}(bz)\big{)}\cos^{2}\big{(}\zeta_{1}^{\varepsilon}(bz)\big{)}\right]dbda. (3.22)

Since RiεR_{i}^{\varepsilon}, i=0,1,,4i=0,1,\cdots,4, are functions depending on the value of ε\varepsilon, the integral with respect to jump measure, IρI_{\rho}, cannot necessarily be expanded a finite linear combination of terms of size εγ\varepsilon^{\gamma} for various γ\gamma.

We thus need the following assumptions on integrability and growth estimates.

  1. (A4)

    For each sufficiently small ε>0\varepsilon>0, the functions logH\log\|\nabla H\| is integrable with respect to μMε\mu_{M}^{\varepsilon}; The functions

    A(x)sinθcosθ,k=1dP~ik(x,θ),i=1,2,3,4,k=1dPjk(x,θ),j=1,2,3 and Iρε(x,θ)A(x)\sin\theta\cos\theta,\;\;\sum_{k=1}^{d}\tilde{P}_{i}^{k}(x,\theta),i=1,2,3,4,\;\;\sum_{k=1}^{d}{P}_{j}^{k}(x,\theta),j=1,2,3\text{ and }I_{\rho}^{\varepsilon}(x,\theta)

    are all integrable with respect to με\mu^{\varepsilon}; There is a constant C<C<\infty such that

    |P~ik(x,θ)με(dx,dθ)|C, for i=2,3,4.\left|\int\tilde{P}_{i}^{k}(x,\theta)\mu^{\varepsilon}(dx,d\theta)\right|\leqslant C,\text{ for }i=2,3,4.
  2. (A5)

    For each sufficiently small ε>0\varepsilon>0, the functions Riε(z)R_{i}^{\varepsilon}(z), i=0,1,,4,i=0,1,\cdots,4, are all integrable with respect to ν\nu, and there is a constant C<C^{\prime}<\infty such that

    |z<cRiε(z)(x,θ)ν(dz)με(dx,dθ)|C, for i=0,1,,4.\left|\int\int_{\|z\|<c}R_{i}^{\varepsilon}(z)(x,\theta)\nu(dz)\mu^{\varepsilon}(dx,d\theta)\right|\leqslant C^{\prime},\text{ for }i=0,1,\cdots,4.

Under assumption (A4)-(A5), by (3.3) and (3.21), it makes sense for us to consider β=23\beta=\frac{2}{3} here. We note that this choice of β\beta is predicated on the fact that the Lévy perturbation in (2.2) consists of a Brownian part and a bounded jump part. To authors’ knowledge, there is no standard value of β\beta in transformation (3.9) for the general Lévy case so far. It is remarkable that there is some flexibility regarding choice of β\beta as we can always get the corresponding estimate results with slight modifications if this parameter takes different values.

We now have the following theorem on estimating the Lyapunov exponent.

Theorem 3.3.

Let β=2/3\beta={2}/{3}. If assumptions (A1-A5) hold, then the Lyapunov exponent for the linearized version (3.2) of the Hamiltonian system with Lévy perturbation (2.2) satisfies

λε=ε23M×S1Σ0(x,θ)με(dx,dθ)+𝒪(ε43)\displaystyle\lambda_{\varepsilon}=\varepsilon^{\frac{2}{3}}\int_{M\times S^{1}}\Sigma_{0}(x,\theta)\mu^{\varepsilon}(dx,d\theta)+\mathcal{O}(\varepsilon^{\frac{4}{3}}) (3.23)

as ε0\varepsilon\to 0, where

Σ0(x,θ)=\displaystyle\Sigma_{0}(x,\theta)= A(x)sinθcosθ+k=1dDk2(x)(12cos2θsin2θcos2θ)+z<cR0ε(z)(x,θ)ν(dz).\displaystyle A(x)\sin\theta\cos\theta+\sum_{k=1}^{d}D_{k}^{2}(x)(\frac{1}{2}\cos^{2}\theta-\sin^{2}\theta\cos^{2}\theta)+\int_{\|z\|<c}R_{0}^{\varepsilon}(z)(x,\theta)\nu(dz). (3.24)
Proof.

We first claim that, for each ε\varepsilon, the processes vtv_{t} and wtw_{t} have the same Lyapunov exponent. Indeed, moving frame (3.5) yields the inequity

min(H,H1)wtvtmax(H,H1)wt.\min(\|\nabla H\|,\|\nabla H\|^{-1})\|w_{t}\|\leqslant\|v_{t}\|\leqslant\max(\|\nabla H\|,\|\nabla H\|^{-1})\|w_{t}\|.

Noting that H\|\nabla H\| is tempered [6, Section 4.1], we have limt1t[logvtlogwt]=0\lim_{t\to\infty}\frac{1}{t}[\log\|v_{t}\|-\log\|w_{t}\|]=0.

By equation (3.18), the Lyapunov exponent of (3.2) is

λε=limt1tlogwt=limt1tρt=limt1t(logw0+0tPε(xs,θs)𝑑s+t1+t2)\displaystyle\lambda_{\varepsilon}=\lim_{t\to\infty}\frac{1}{t}\log\|w_{t}\|=\lim_{t\to\infty}\frac{1}{t}\rho_{t}=\lim_{t\to\infty}\frac{1}{t}\left(\log\|w_{0}\|+\int_{0}^{t}{P}_{\varepsilon}(x_{s},\theta_{s})ds+\mathcal{M}_{t}^{1}+\mathcal{M}_{t}^{2}\right) (3.25)

with

Pε(x,θ)=ε23A(x)sinθcosθ+σ~k2(x,θ)+Iρε(x,θ),{P}_{\varepsilon}(x,\theta)=\varepsilon^{\frac{2}{3}}A(x)\sin\theta\cos\theta+\tilde{\sigma}_{k}^{2}(x,\theta)+I_{\rho}^{\varepsilon}(x,\theta),
t1=0tk=1dσk2(xs,θs)dBsk and t2=0tz<c[ζ2ε(z)(xs,θs)]N~(ds,dz).\mathcal{M}_{t}^{1}=\int_{0}^{t}\sum_{k=1}^{d}\sigma_{k}^{2}(x_{s},\theta_{s})dB_{s}^{k}\;\text{ and }\;\mathcal{M}_{t}^{2}=\int_{0}^{t}\int_{\|z\|<c}\left[\zeta_{2}^{\varepsilon}(z)(x_{s-},\theta_{s-})\right]\tilde{N}(ds,dz).

Referring to [12, Page 72] and [4, Page 109], both t1\mathcal{M}_{t}^{1} and t2\mathcal{M}_{t}^{2} are martingales for which ti/t0\mathcal{M}_{t}^{i}/t\to 0, i=1,2i=1,2, as tt\to\infty. The ergodic theorem and assumption (A3) yields the Khas’minskii formula

λε=M×S1Pε(x,θ)𝑑ε(x,θ).\displaystyle\lambda_{\varepsilon}=\int_{M\times S^{1}}P_{\varepsilon}(x,\theta)d\mathbb{P}_{\varepsilon}(x,\theta). (3.26)

The result now follows easily by using assumptions (A4) and (A5). ∎

Remark 3.4.

Theorem 3.3 shows that, to estimate the Lyapunov exponent for the original linearization system (3.2), we could simplify the problem by analyzing the leading drift and diffusion terms of the transformed system (3.17)-(3.18) in the sense of a Pinsky-Wihstutz transformation (3.9). This result is essentially based on the fact that the linearization of a one degree of freedom Hamiltonian system, when written with respect to a suitable moving frame, is a nilpotent linear system.

Remark 3.5.

The result in Theorem 3.3 depends closely on the type of perturbations which is indeed a combination of Brownian part and a bounded jump part here. If there is no jump term, then the result in agreement with that given in [10]. If the Brownian term is vanished and the assumption (A5) is failed, then the result may need to be modified with respect to other value of β\beta. We also note that the assumption (A5) is almost harsh, even though it could be satisfied in some simple case (see Example 4.1). Without this assumption, Theorem 3.3 still makes sense if we replace the third term of Σ0\Sigma_{0} by ε23Iρε\varepsilon^{-\frac{2}{3}}I_{\rho}^{\varepsilon} with IρεI_{\rho}^{\varepsilon} given in (3.21).

Remark 3.6.

Here we only focus on the theoretical expression of the Lyapunov exponent, and the main novelty of our work is the model itself. We have to point out that it is still a challenging but important task to develop efficient methods for estimating the integral in (3.23), which is left for our future research. Actually, for Brownian case, there is a remarkable method to deal with this problem by combining averaging arguments and an adjoint expansion method in non-compact spaces; see [10] for details. However, note that the stochastic averaging theories for SDE with multiplicative Lévy noise is still remain to solve and the corresponding generator is indeed a non-local operator, this method is unfit to the Lévy case directly. For some interesting works on evaluating numerical values of integrals with respect to invariant measures, we refer to [5, 2, 24] and the references therein.

4 Examples

In this section, we present two simple illustrative examples of the theory developed in the previous section.

Example 4.1.

(A stochastic nilpotent linear system) We first recall that a symmetric α\alpha-stable Lévy motion LtαL_{t}^{\alpha} is a special Lévy motion, with non-Gaussianity index (or stability index) α(0,2)\alpha\in(0,2). It has the generating triplet (0,0,να)(0,0,\nu_{\alpha}) with jump measure να=Cαdu|u|d+α\nu_{\alpha}=C_{\alpha}\frac{du}{|u|^{d+\alpha}} [12].

Consider the following linear SDE in 2\{0}\mathbb{R}^{2}\backslash\{0\}:

dut=[0a00]utdt+ε[00σ0]utdBt+ε[00σ0]utdLtα,\displaystyle du_{t}=\begin{bmatrix}\begin{matrix}0&a\\ 0&0\end{matrix}\end{bmatrix}u_{t}dt+\varepsilon\begin{bmatrix}\begin{matrix}0&0\\ \sigma&0\end{matrix}\end{bmatrix}u_{t}\circ dB_{t}+\varepsilon\begin{bmatrix}\begin{matrix}0&0\\ \sigma&0\end{matrix}\end{bmatrix}u_{t}\diamond dL^{\alpha}_{t}, (4.1)

which can be regarded as a perturbed Hamiltonian system with H(u1,u2)=12au22H(u_{1},u_{2})=\frac{1}{2}au_{2}^{2}. Here a>0a>0, σ>0\sigma>0 and LtαL^{\alpha}_{t} (1<α<2)(1<\alpha<2) is a one-dimensional standard α\alpha-stable Lévy motion. Note that, for this system, the time one flow is ξ(z)(u)=(u1,zσu1+u2)\xi(z)(u)=(u_{1},z\sigma u_{1}+u_{2}) and the jump term 0st[ξ(ΔLsα)(us)us(0,us1ΔLsα)]=0\sum_{0\leqslant s\leqslant t}\left[\xi(\Delta L^{\alpha}_{s})(u_{s-})-u_{s-}-(0,u_{s-}^{1}\Delta L^{\alpha}_{s})\right]=0, so that the coefficients of both Marcus and Itô versions are identical.

By introducing the following change of variables : (θ(0,2π)\{π2,3π2})(\theta\in(0,2\pi)\backslash\{\frac{\pi}{2},\frac{3\pi}{2}\})

cosθ=u1/ut,sinθ=u2/u,ρ=logu,\cos\theta=u_{1}/\|u_{t}\|,\;\;\;\;\;\sin\theta=u_{2}/\|u\|,\;\;\;\;\;\rho=\log\|u\|,

we have

d[θtρt]=a[sin2θtsinθtcosθt]dt+εσ[cos2θtsinθtcosθt]dBt+εσ[cos2θtsinθtcosθt]dLtα,\displaystyle d\begin{bmatrix}\begin{matrix}\theta_{t}\\ \rho_{t}\end{matrix}\end{bmatrix}=a\begin{bmatrix}\begin{matrix}-\sin^{2}\theta_{t}\\ \sin\theta_{t}\cos\theta_{t}\end{matrix}\end{bmatrix}dt+\varepsilon\sigma\begin{bmatrix}\begin{matrix}\cos^{2}\theta_{t}\\ \sin\theta_{t}\cos\theta_{t}\end{matrix}\end{bmatrix}\circ dB_{t}+\varepsilon\sigma\begin{bmatrix}\begin{matrix}\cos^{2}\theta_{t}\\ \sin\theta_{t}\cos\theta_{t}\end{matrix}\end{bmatrix}\diamond dL^{\alpha}_{t}, (4.2)

or equivalently,

dθt=\displaystyle d\theta_{t}= asin2θtdtε2σ2sinθtcos3θtdt+εσcos2θtdBt\displaystyle-a\sin^{2}\theta_{t}dt-\varepsilon^{2}\sigma^{2}\sin\theta_{t}\cos^{3}\theta_{t}dt+\varepsilon\sigma\cos^{2}\theta_{t}dB_{t}
+|z|<c[ζ1(z)(θt)θt]να(dz)𝑑t\displaystyle+\int_{|z|<c}[\zeta_{1}(z)(\theta_{t-})-\theta_{t-}]\nu_{\alpha}(dz)dt
+|z|<c[ζ1(z)(θt)θt]N~(dt,dz)+|z|c[ζ1(z)(θt)θt]N(dt,dz),\displaystyle+\int_{|z|<c}[\zeta_{1}(z)(\theta_{t-})-\theta_{t-}]\tilde{N}(dt,dz)+\int_{|z|\geqslant c}[\zeta_{1}(z)(\theta_{t-})-\theta_{t-}]N(dt,dz), (4.3)
dρt=\displaystyle d\rho_{t}= asinθtcosθtdt+ε2σ2(12cos2θtsin2θtcos2θt)dt+εσsinθtcosθtdBt\displaystyle a\sin\theta_{t}\cos\theta_{t}dt+\varepsilon^{2}\sigma^{2}\Big{(}\frac{1}{2}\cos^{2}\theta_{t}-\sin^{2}\theta_{t}\cos^{2}\theta_{t}\Big{)}dt+\varepsilon\sigma\sin\theta_{t}\cos\theta_{t}dB_{t}
+|z|<c[ζ2(z)(θt)]να(dz)𝑑t\displaystyle+\int_{|z|<c}[\zeta_{2}(z)(\theta_{t-})]\nu_{\alpha}(dz)dt
+|z|<c[ζ2(z)(θt)]N~(dt,dz)+|z|c[ζ2(z)(θt)]N(dt,dz),\displaystyle+\int_{|z|<c}[\zeta_{2}(z)(\theta_{t-})]\tilde{N}(dt,dz)+\int_{|z|\geqslant c}[\zeta_{2}(z)(\theta_{t-})]N(dt,dz), (4.4)

with

ζ1(z)(θ)\displaystyle\zeta_{1}(z)(\theta) =arctan(tanθ+εσz),\displaystyle=\arctan(\tan\theta+\varepsilon\sigma z), (4.5)
ζ2(z)(θ)\displaystyle\zeta_{2}(z)(\theta) =12log(1+(tanθ+εσz)21+tan2θ).\displaystyle=\frac{1}{2}\log\left(\frac{1+(\tan\theta+\varepsilon\sigma z)^{2}}{1+\tan^{2}\theta}\right). (4.6)

By interlacing technique, we consider the modified version of equations (4.3)-(4.4). Provided the modified process ρ(t)\rho(t) is ergodic, the Lyapunov exponent can be obtained as

λ=limt1tlogut=limt1tρt=𝔼[Q(θ)]=02πQ(θ)μ(θ)𝑑θ,\displaystyle\lambda=\lim_{t\to\infty}\frac{1}{t}\log\|u_{t}\|=\lim_{t\to\infty}\frac{1}{t}\rho_{t}=\mathbb{E}[Q(\theta)]=\int_{0}^{2\pi}Q(\theta)\mu(\theta)d\theta, (4.7)

where

Q(θ)=\displaystyle Q(\theta)= asinθcosθ+ε2σ2(12cos2θtsin2θtcos2θt)+|z|<c[ζ2(z)(θ)]να(dz),\displaystyle a\sin\theta\cos\theta+\varepsilon^{2}\sigma^{2}\Big{(}\frac{1}{2}\cos^{2}\theta_{t}-\sin^{2}\theta_{t}\cos^{2}\theta_{t}\Big{)}+\int_{|z|<c}[\zeta_{2}(z)(\theta)]\nu_{\alpha}(dz), (4.8)

and μ(θ)\mu(\theta) is the density of the invariant measure of the modified process θt\theta_{t} with respect to the uniform measure on the unit circle. Remark that the operator for modified process θt\theta_{t} satisfies

𝒜θf(θ)=\displaystyle\mathcal{A}_{\theta}f(\theta)= (asin2θθ)f(θ)+ε2σ2(sinθcos3θθ+12cos4θ2θ2)f(θ)\displaystyle-(a\sin^{2}\theta\frac{\partial}{\partial\theta})f(\theta)+\varepsilon^{2}\sigma^{2}\Big{(}-\sin\theta\cos^{3}\theta\frac{\partial}{\partial\theta}+\frac{1}{2}\cos^{4}\theta\frac{\partial^{2}}{\partial\theta^{2}}\Big{)}f(\theta)
+|z|<c[f(ζ1(z)(θ))f(θ)]να(dz).\displaystyle+\int_{|z|<c}\Big{[}f(\zeta_{1}(z)(\theta))-f(\theta)\Big{]}\nu_{\alpha}(dz). (4.9)

Referring to Sun et. al. [22], μ(θ)\mu(\theta) is indeed the solution of the following stationary Fokker-Planck equation

𝒜θμ(θ)=0,\displaystyle\mathcal{A}_{\theta}^{\ast}\mu(\theta)=0, (4.10)

that is,

cos2θθ(sin2θμ(θ))+12ε2σ2cos2θθ(cos2θθ(cos2θμ(θ)))\displaystyle\cos^{2}\theta\frac{\partial}{\partial\theta}(\sin^{2}\theta\mu(\theta))+\frac{1}{2}\varepsilon^{2}\sigma^{2}\cos^{2}\theta\frac{\partial}{\partial\theta}\left(\cos^{2}\theta\frac{\partial}{\partial\theta}(\cos^{2}\theta\mu(\theta))\right)
+\{0}[cos2(ζ1(z)(θ))μ(ζ1(z)(θ))cos2θμ(θ)]να(dz)=0.\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+\int_{\mathbb{R}\backslash\{0\}}\left[\cos^{2}(\zeta_{1}(z)(\theta))\mu(\zeta_{1}(z)(\theta))-\cos^{2}\theta\mu(\theta)\right]\nu_{\alpha}(dz)=0. (4.11)

Notice that, by Taylor’s Theorem, we have

|z|<c[ζ2(z)(θ)]να(dz)=\displaystyle\int_{|z|<c}[\zeta_{2}(z)(\theta)]\nu_{\alpha}(dz)= |z|<c[g(εσz)(θ)g(0)(θ)]Cα|z|1+α𝑑z\displaystyle\int_{|z|<c}[g(\varepsilon\sigma z)(\theta)-g(0)(\theta)]\frac{C_{\alpha}}{|z|^{1+\alpha}}dz
=\displaystyle= |z|<c[12g′′(0)(θ)ε2σ2z2+14!01g(4)(τεσz)(θ)ε4σ4z4𝑑τ]Cα|z|1+α𝑑z,\displaystyle\int_{|z|<c}\left[\frac{1}{2}g^{\prime\prime}(0)(\theta)\varepsilon^{2}\sigma^{2}z^{2}+\frac{1}{4!}\int_{0}^{1}g^{(4)}(\tau\varepsilon\sigma z)(\theta)\varepsilon^{4}\sigma^{4}z^{4}d\tau\right]\frac{C_{\alpha}}{|z|^{1+\alpha}}dz,

where g()(θ)=12log(1+(tanθ+)2)g(\cdot)(\theta)=\frac{1}{2}\log(1+(\tan\theta+\cdot)^{2}), g(0)=tanθ1+tan2θ=sinθcosθg^{\prime}(0)=\frac{\tan\theta}{1+\tan^{2}\theta}=\sin\theta\cos\theta, 12g′′(0)=121tan2θ(1+tan2θ)2=12cos2θsin2θcos2θ\frac{1}{2}g^{\prime\prime}(0)=\frac{1}{2}\frac{1-\tan^{2}\theta}{(1+\tan^{2}\theta)^{2}}=\frac{1}{2}\cos^{2}\theta-\sin^{2}\theta\cos^{2}\theta and g(4)(σz)=2(tanθ+σz)4+(tanθ+σz)2+6(tanθ+σz)1[1+(tanθ+σz)2]4g^{(4)}(\sigma z)=\frac{-2(\tan\theta+\sigma z)^{4}+(\tan\theta+\sigma z)^{2}+6(\tan\theta+\sigma z)-1}{[1+(\tan\theta+\sigma z)^{2}]^{4}}. It is not hard to show that, for all θ(π2,π2)(π2,3π2)\theta\in(-\frac{\pi}{2},\frac{\pi}{2})\bigcup(\frac{\pi}{2},\frac{3\pi}{2}) and ε(0,1)\varepsilon\in(0,1), there is a positive constant C1C_{1} such that

14!|z|<c[g(4)(δεσz)(θ)σ4z4]Cα|z|1+α𝑑z<C1.\displaystyle\frac{1}{4!}\int_{|z|<c}\left[g^{(4)}(\delta\varepsilon\sigma z)(\theta)\sigma^{4}z^{4}\right]\frac{C_{\alpha}}{|z|^{1+\alpha}}dz<C_{1}.

Analogously, we have

|z|<c[f(ζ1(z)(θ))f(θ)]να(dz)=|z|<c[f(ζ1(z)(θ))f(ζ1(0)(θ))]Cα|z|1+α𝑑z\displaystyle\int_{|z|<c}\Big{[}f(\zeta_{1}(z)(\theta))-f(\theta)\Big{]}\nu_{\alpha}(dz)=\int_{|z|<c}\Big{[}f(\zeta_{1}(z)(\theta))-f(\zeta_{1}(0)(\theta))\Big{]}\frac{C_{\alpha}}{|z|^{1+\alpha}}dz
=|z|<c[12F′′(0)(θ)ε2σ2z2+14!01F(4)(τεσz)(θ)ε4σ4z4𝑑τ]Cα|z|1+α𝑑z,\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}=\int_{|z|<c}\left[\frac{1}{2}F^{\prime\prime}(0)(\theta)\varepsilon^{2}\sigma^{2}z^{2}+\frac{1}{4!}\int_{0}^{1}F^{(4)}(\tau\varepsilon\sigma z)(\theta)\varepsilon^{4}\sigma^{4}z^{4}d\tau\right]\frac{C_{\alpha}}{|z|^{1+\alpha}}dz,

where F()(θ)=farctan(tanθ+)F(\cdot)(\theta)=f\circ\arctan(\tan\theta+\cdot) and 12F′′(0)=tanθ(1+tan2θ)2f(θ)+121(1+tan2θ)2f′′(θ)=(cos3θsinθθ+12cos4θ2θ2)f(θ)\frac{1}{2}F^{\prime\prime}(0)=-\frac{\tan\theta}{(1+\tan^{2}\theta)^{2}}f^{\prime}(\theta)+\frac{1}{2}\frac{1}{(1+\tan^{2}\theta)^{2}}f^{\prime\prime}(\theta)=\big{(}-\cos^{3}\theta\sin\theta\frac{\partial}{\partial\theta}+\frac{1}{2}\cos^{4}\theta\frac{\partial^{2}}{\partial\theta^{2}}\big{)}f(\theta).

By taking Pinsky-Wihstutz transformation (3.9), u~t=Tut\tilde{u}_{t}=Tu_{t}, with β=23\beta=\frac{2}{3}, and introducing the corresponding change of variables with respect to (θ~t,ρ~t)(\tilde{\theta}_{t},\tilde{\rho}_{t}), we conclude that

λ=limt1tlogut=limt1tlogTut=limt1tρ~t.\displaystyle\lambda=\lim_{t\to\infty}\frac{1}{t}\log\|u_{t}\|=\lim_{t\to\infty}\frac{1}{t}\log\|Tu_{t}\|=\lim_{t\to\infty}\frac{1}{t}\tilde{\rho}_{t}. (4.12)

According to Theorem 3.3, we further have

λ=ε23S1Q~(θ~)μ~(θ~)𝑑θ~+𝒪(ε43),\displaystyle\lambda=\varepsilon^{\frac{2}{3}}\int_{S^{1}}\tilde{Q}(\tilde{\theta})\tilde{\mu}(\tilde{\theta})d\tilde{\theta}+\mathcal{O}(\varepsilon^{\frac{4}{3}}), (4.13)

where

Q~(θ)=\displaystyle\tilde{Q}(\theta)= asinθcosθ+σ2(12cos2θsin2θcos2θ)(1+|z|<cz2να(dz))\displaystyle a\sin\theta\cos\theta+\sigma^{2}\Big{(}\frac{1}{2}\cos^{2}\theta-\sin^{2}\theta\cos^{2}\theta\Big{)}\Big{(}1+\int_{|z|<c}z^{2}\nu_{\alpha}(dz)\Big{)}
=\displaystyle= asinθcosθ+σ2(1+2Cα2αc2α)(12cos2θsin2θcos2θ),\displaystyle a\sin\theta\cos\theta+\sigma^{2}\Big{(}1+\frac{2C_{\alpha}}{2-\alpha}c^{2-\alpha}\Big{)}\Big{(}\frac{1}{2}\cos^{2}\theta-\sin^{2}\theta\cos^{2}\theta\Big{)}, (4.14)

and μ~(θ)\tilde{\mu}(\theta) is the corresponding density of the invariant measure of the process θ~t\tilde{\theta}_{t}. Moreover, define

𝒜~θ1=(asin2θθ)+σ2(1+2Cα2αc2α)(sinθcos3θθ+12cos4θ2θ2).\displaystyle\tilde{\mathcal{A}}^{1}_{\theta}=-(a\sin^{2}\theta\frac{\partial}{\partial\theta})+\sigma^{2}\Big{(}1+\frac{2C_{\alpha}}{2-\alpha}c^{2-\alpha}\Big{)}\Big{(}-\sin\theta\cos^{3}\theta\frac{\partial}{\partial\theta}+\frac{1}{2}\cos^{4}\theta\frac{\partial^{2}}{\partial\theta^{2}}\Big{)}. (4.15)

Then the generator for the process θ~t\tilde{\theta}_{t} is

𝒜~θ=\displaystyle\tilde{\mathcal{A}}_{\theta}= ε23𝒜~θ1+ε4314!σ4|z|<c01F(4)(τεσz)(θ)z4𝑑τCα|z|1+α𝑑z.\displaystyle\varepsilon^{\frac{2}{3}}\tilde{\mathcal{A}}^{1}_{\theta}+\varepsilon^{\frac{4}{3}}\cdot\frac{1}{4!}\sigma^{4}\int_{|z|<c}\int_{0}^{1}F^{(4)}(\tau\varepsilon\sigma z)(\theta)z^{4}d\tau\frac{C_{\alpha}}{|z|^{1+\alpha}}dz. (4.16)
Example 4.2.

(A stochastic Duffing equation) This is a Lévy version of (last) example of [10]. We will show that the viewpoint of introducing a suitable moving frame to converting the original linearization system to a nilpotent one, and our estimation method method based on transformation (3.9) is applicable in the present example.

Consider a stochastic Lévy version of the Duffing equation in 2\{0}\mathbb{R}^{2}\backslash\{0\}:

x¨=xx3+εσxdLt\displaystyle\ddot{x}=-x-x^{3}+\varepsilon\sigma x\diamond dL_{t} (4.17)

with σ>0\sigma>0 and Ltα(0,I,ν)L^{\alpha}_{t}\sim(0,I,\nu) a one-dimensional Lévy motion. Rewriting (4.17) in phase space (x,y)=(x,x˙)2(x,y)=(x,\dot{x})\in\mathbb{R}^{2}, we have

U1(x,y)=[yxx3],U2(x,y)=1(x+x3)2+y2[x+x3y].U_{1}(x,y)=\begin{bmatrix}\begin{matrix}y\\ -x-x^{3}\end{matrix}\end{bmatrix},\;\;\;U_{2}(x,y)=\frac{1}{(x+x^{3})^{2}+y^{2}}\begin{bmatrix}\begin{matrix}x+x^{3}\\ y\end{matrix}\end{bmatrix}.

This system (4.17) is indeed a perturbation of the Hamiltonian system with H(x,y)=12x2+14x4+12y2H(x,y)=\frac{1}{2}x^{2}+\frac{1}{4}x^{4}+\frac{1}{2}y^{2} by the following vector fields

V1(x,y)=σ[0x].V_{1}(x,y)=\sigma\begin{bmatrix}\begin{matrix}0\\ x\end{matrix}\end{bmatrix}.

Linearizing system (4.17) along the trajectory (xt,yt)(x_{t},y_{t}) and introducing a moving frame in the form of (3.3) with a11(x,y)=εσx(x+x3)(x+x3)2+y2a_{1}^{1}(x,y)=-\frac{\varepsilon\sigma x(x+x^{3})}{(x+x^{3})^{2}+y^{2}} and a12(x,y)=σxya_{1}^{2}(x,y)=\sigma xy, we obtain

dwt=[0A00]wtdt+ε[B1C1D1E1]wtdLt,\displaystyle dw_{t}=\begin{bmatrix}\begin{matrix}0&A\\ 0&0\end{matrix}\end{bmatrix}w_{t}dt+\varepsilon\begin{bmatrix}\begin{matrix}B_{1}&C_{1}\\ D_{1}&E_{1}\end{matrix}\end{bmatrix}w_{t}\diamond dL_{t}, (4.18)

where A=3x2((x+x3)2y2)[(x+x3)2+y2]2A=\frac{3x^{2}((x+x^{3})^{2}-y^{2})}{[(x+x^{3})^{2}+y^{2}]^{2}}, B1=E1=σxy(x2+2)(x+x3)2+y2B_{1}=-E_{1}=\frac{-\sigma xy(x^{2}+2)}{(x+x^{3})^{2}+y^{2}}, C1=σx3(x+x3)[(x+x3)2+y2]2C_{1}=\frac{-\sigma x^{3}(x+x^{3})}{[(x+x^{3})^{2}+y^{2}]^{2}}, D1=σx(x+x3)+σy2D_{1}=-\sigma x(x+x^{3})+\sigma y^{2}.

Let β=23\beta=\frac{2}{3}, we further introduce Pinsky-Wihstutz transform (3.9) and rewrite the transformed system with respect to w=w[cosθ,sinθ]Tw=\|w\|[\cos\theta,\sin\theta]^{T} and ρ=logw\rho=\log\|w\|:

dθt=\displaystyle d\theta_{t}= ε23A(xt,yt)sin2θt+σ11(xt,yt,θt)dLtk,\displaystyle-\varepsilon^{\frac{2}{3}}A(x_{t},y_{t})\sin^{2}\theta_{t}+\sigma_{1}^{1}(x_{t},y_{t},\theta_{t})\diamond dL^{k}_{t}, (4.19)
dρt=\displaystyle d\rho_{t}= ε23A(xt,yt)sinθtcosθt+σ12(xt,yt,θt)dLtk,\displaystyle\varepsilon^{\frac{2}{3}}A(x_{t},y_{t})\sin\theta_{t}\cos\theta_{t}+\sigma_{1}^{2}(x_{t},y_{t},\theta_{t})\diamond dL^{k}_{t}, (4.20)

where

σ11(u,θ)=\displaystyle\sigma_{1}^{1}(u,\theta)= ε13D1(x,y)cos2θεB1(x,y)sin2θε53C1(x,y)sin2θ,\displaystyle\varepsilon^{\frac{1}{3}}D_{1}(x,y)\cos^{2}\theta-\varepsilon B_{1}(x,y)\sin 2\theta-\varepsilon^{\frac{5}{3}}C_{1}(x,y)\sin^{2}\theta,
σ12(u,θ)=\displaystyle\sigma_{1}^{2}(u,\theta)= ε13D1(x,y)sinθcosθ+εB1(x,y)cos2θ+ε53C1(x,y)sinθcosθ.\displaystyle\varepsilon^{\frac{1}{3}}D_{1}(x,y)\sin\theta\cos\theta+\varepsilon B_{1}(x,y)\cos 2\theta+\varepsilon^{\frac{5}{3}}C_{1}(x,y)\sin\theta\cos\theta.

We next focus on the modified linearization system of (4.17) in the sense of interlacing. By Theorem 3.3 and Remark 3.5, the Lyapunov exponent of the modified linearization system is

λ=ε232\{0}×S1Σ0(x,y,θ)με(dx,dy,dθ)+𝒪(ε43),\displaystyle\lambda=\varepsilon^{\frac{2}{3}}\int_{\mathbb{R}^{2}\backslash\{0\}\times S^{1}}\Sigma_{0}(x,y,\theta)\mu^{\varepsilon}(dx,dy,d\theta)+\mathcal{O}(\varepsilon^{\frac{4}{3}}), (4.21)

where

Σ0(x,y,θ)=\displaystyle\Sigma_{0}(x,y,\theta)= A(x,y)sinθcosθ+D12(x,y)(12cos2θsin2θcos2θ)\displaystyle A(x,y)\sin\theta\cos\theta+D_{1}^{2}(x,y)(\frac{1}{2}\cos^{2}\theta-\sin^{2}\theta\cos^{2}\theta)
+ε23z<c[ζ2(z)(x,θ)zσ12(x,θ)]ν(dz)\displaystyle+\varepsilon^{-\frac{2}{3}}\int_{\|z\|<c}\left[\zeta_{2}(z)(x,\theta)-z\sigma_{1}^{2}(x,\theta)\right]\nu(dz) (4.22)

with ζ2\zeta_{2} the time one flow of the vector field σ12\sigma_{1}^{2}. According to the analysis of (3.21), the leading order of the third term in (4.2) should not less than ε0=1\varepsilon^{0}=1. But unfortunately, it is not easy to obtain the precise expression of ζ2\zeta_{2}, even though we know that ξ(τz)(u)=(x,εστzx+y)T\xi(\tau z)(u)=(x,\varepsilon\sigma\tau zx+y)^{T}. Compared with the Brownian case in [10], an extra term (the third one in (4.2)) appears in the integrand of integral expression for Lyapunov exponent. If the modified system of (4.17) is only driven by Brownian noise, that is, ν=0\nu=0, then the result here would be consistent with that in [10].

Data Availability

The data that support the findings of this study are available within the article.

Acknowledgements

The authors would like to thank Dr Qiao Huang, Dr Li Lin, Dr Zibo Wang, and Dr Yanjie Zhang for helpful discussions. This work was partly supported by NSFC grants 11771449 and 11531006.

References

References

  • [1] S. Albeverio, Z. Brzeźniak, and J-L. Wu. Existence of global solutions and invariant measures for stochastic differential equations driven by Poisson type noise with non-Lipschitz coefficients. Journal of Mathematical Analysis and Applications, 371(1):309–322, 2010.
  • [2] S. Albeverio, L. D. Persio, E. Mastrogiacomo, and B. Smii. A class of Lévy driven SDEs and their explicit invariant measures. Potential Analysis, 45(2):229–259, 2014.
  • [3] S. Albeverio, B. Rüdiger, and J-L. Wu. Invariant Measures and symmetry property of Lévy type operators. Potential Analysis, 13(2):147–168, 2000.
  • [4] D. Applebaum. Lévy Processes and Stochastic Calculus. Cambridge University Press, 2009.
  • [5] S. T. Ariaratnam and W. C. Xie. Lyapunov exponent and rotation number of a two-dimensional nilpotent stochastic system. Dynamics and Stability of Systems, 5(1):1–9, 1990.
  • [6] L. Arnold. Random Dynamical Systems. Springer, 1998.
  • [7] L. Arnold, E. Oeljeklaus, and E. Pardoux. Almost sure and moment stability for linear Itô equations. In Lyapunov Exponents, Lecture Notes Mathematics, pages 129–159. Springer, 1986.
  • [8] P. H. Baxendale. Stability along trajectories at a stochastic bifurcation point. In Stochastic Dynamic, pages 1–25. Springer, Berlin, 1999.
  • [9] P. H. Baxendale and L. Goukasian. Lyapunov exponents of nilpotent Itô systems with random coefficients. Stochastic Processes and their Applications, 95:219–233, 2001.
  • [10] P. H. Baxendale and L. Goukasian. Lyapunov exponents for small random perturbations of Hamiltonian systems. Annals of Probability, 30(1):101–134, 2002.
  • [11] J-M. Bismut. Mécanique Aléatoire, Lecture Notes in Mathematics, volume 866. Springer-Verlag, 1981.
  • [12] J. Duan. An Introduction to Stochastic Dynamics. Cambridge University Press, 2015.
  • [13] M. I. Freidlin and A. D. Wentzell. Random Perturbations of Dynamical Systems, Third Edition. Springer-Verlag, New York, 2012.
  • [14] R. Z. Khasminskii. Necessary and sufficient conditions for the asymptotic stability of linear stochastic systems. Theory of Probability & Its Applications, 12:144–147, 1967.
  • [15] T. G. Kurtz, E. Pardoux, and P. Protter. Stratonovich stochastic differential equations driven by general semimartingales. In Annales de l’IHP Probabilités et statistiques, volume 31, pages 351–377, 1995.
  • [16] X. Mao. Stochastic Differential Equations and Applications, Second Edition. Horwood Publishing, 2006.
  • [17] X. Mao and A. E. Rodkina. Exponential stability of stochastic differential equations driven by discontinuous semimartingales. Stochastics & Stochastic Reports, 55:207–224, 1995.
  • [18] S. I. Marcus. Modeling and approximation of stochastic differential equations driven by semimartingales. Stochastics: An International Journal of Probability and Stochastic Processes, 4(3):223–245, 1981.
  • [19] M. A. Pinsky and V. Wihstutz. Lyapunov exponents of nilpotent ito systems. Stochastics: formerly Stochastics and Stochastics Reports, 25(1):43–57, 1988.
  • [20] H. Qiao and J. Duan. Lyapunov exponents of stochastic differential equations driven by Lévy processes. Dynamical Systems, 31:136–150, 2016.
  • [21] K-I Sato. Lévy Processes and Infinitely Divisible Distributions. Cambridge University Press, 1999.
  • [22] X. Sun, J. Duan, X. Li, H. Liu, X. Wang, and Y. Zheng. Derivation of Fokker-Planck equations for stochastic dynamical systems under excitation of multiplicative non-Gaussian white noise. Journal of Mathematical Analysis & Applications, 446(1):786–800, 2017.
  • [23] P. Wei, Y. Chao, and J. Duan. Hamiltonian systems with Lévy noise: Symplecticity, Hamilton’s principle and averaging principle. Physica D: Nonlinear Phenomena, 398:69–83, 2019.
  • [24] Y. Zhang, X. Wang, Q. Huang, J. Duan, and T. Li. Numerical analysis and applications of Fokker-Planck equations for stochastic dynamical systems with multiplicative α\alpha-stable noises. Applied Mathematical Modelling, 87:711–730, 2020.
  • [25] W. Zhu. Lyapunov exponent and stochastic stability of quasi-non-integrable Hamiltonian systems. International Journal of Non-Linear Mechanics, 39(4):569–579, 2004.