
In this paper, stability of highly nonlinear hybrid neutral stochastic differential delay equations (NSDDEs) is discussed. In contrast to the white noise examined in previous literature, we incorporate colored noise into the highly nonlinear hybrid NSDDEs. Under some assumptions, we can show that highly nonlinear hybrid NSDDEs have a unique global solution. Meanwhile, we establish some criteria related to noise-to-state stability (NSS) of global solutions. Additionally, some theorems are given to guarantee asymptotic stability in Lˆα and almost surely asymptotic stability of global solution. These related discriminant rules are delay-dependent. Finally, an example is provided to demonstrate the validity of theoretical results.
Citation: Siru Li, Tian Xu, Ailong Wu. Existence and asymptotic properties of global solution for hybrid neutral stochastic differential delay equations with colored noise[J]. AIMS Mathematics, 2025, 10(3): 6379-6405. doi: 10.3934/math.2025291
[1] | Tian Xu, Ailong Wu . Stabilization of nonlinear hybrid stochastic time-delay neural networks with Lévy noise using discrete-time feedback control. AIMS Mathematics, 2024, 9(10): 27080-27101. doi: 10.3934/math.20241317 |
[2] | Jingjing Yang, Jianqiu Lu . Stabilization in distribution of hybrid stochastic differential delay equations with Lévy noise by discrete-time state feedback controls. AIMS Mathematics, 2025, 10(2): 3457-3483. doi: 10.3934/math.2025160 |
[3] | Zhifu Jia, Xinsheng Liu . New stability theorems of uncertain differential equations with time-dependent delay. AIMS Mathematics, 2021, 6(1): 623-642. doi: 10.3934/math.2021038 |
[4] | Leonid Shaikhet . About one unsolved problem in asymptotic p-stability of stochastic systems with delay. AIMS Mathematics, 2024, 9(11): 32571-32577. doi: 10.3934/math.20241560 |
[5] | Yijia Zhang, Tao Xie, Yunlong Ma . Robustness analysis of exponential stability of Cohen-Grossberg neural network with neutral terms. AIMS Mathematics, 2025, 10(3): 4938-4954. doi: 10.3934/math.2025226 |
[6] | Zhengwen Yin, Yuanshun Tan . Threshold dynamics of stochastic SIRSW infectious disease model with multiparameter perturbation. AIMS Mathematics, 2024, 9(12): 33467-33492. doi: 10.3934/math.20241597 |
[7] | Wedad Albalawi, Muhammad Imran Liaqat, Fahim Ud Din, Kottakkaran Sooppy Nisar, Abdel-Haleem Abdel-Aty . Well-posedness and Ulam-Hyers stability results of solutions to pantograph fractional stochastic differential equations in the sense of conformable derivatives. AIMS Mathematics, 2024, 9(5): 12375-12398. doi: 10.3934/math.2024605 |
[8] | Omar Kahouli, Saleh Albadran, Zied Elleuch, Yassine Bouteraa, Abdellatif Ben Makhlouf . Stability results for neutral fractional stochastic differential equations. AIMS Mathematics, 2024, 9(2): 3253-3263. doi: 10.3934/math.2024158 |
[9] | Chunli You, Linxin Shu, Xiao-bao Shu . Approximate controllability of second-order neutral stochastic differential evolution systems with random impulsive effect and state-dependent delay. AIMS Mathematics, 2024, 9(10): 28906-28930. doi: 10.3934/math.20241403 |
[10] | Ping Zhu . Dynamics of the positive almost periodic solution to a class of recruitment delayed model on time scales. AIMS Mathematics, 2023, 8(3): 7292-7309. doi: 10.3934/math.2023367 |
In this paper, stability of highly nonlinear hybrid neutral stochastic differential delay equations (NSDDEs) is discussed. In contrast to the white noise examined in previous literature, we incorporate colored noise into the highly nonlinear hybrid NSDDEs. Under some assumptions, we can show that highly nonlinear hybrid NSDDEs have a unique global solution. Meanwhile, we establish some criteria related to noise-to-state stability (NSS) of global solutions. Additionally, some theorems are given to guarantee asymptotic stability in Lˆα and almost surely asymptotic stability of global solution. These related discriminant rules are delay-dependent. Finally, an example is provided to demonstrate the validity of theoretical results.
Numerous stochastic dynamical systems demonstrate dependencies on both current and previous states, while also integrating delayed derivatives. In order to more accurately describe and simulate such systems, neutral stochastic differential equations are commonly employed [1]. In practical applications, the time delay effect is a critical factor in characterizing the dynamical behavior of systems [2]. For instance, synaptic signal transmission in biological neural networks involves axonal conduction delay. At the same time, communication delay in industrial networked control systems also requires modeling through a delay term [3], such as W(Φ(t−δ)). Neutral stochastic differential delay equations (NSDDEs) with Markov switching constitute a significant class of hybrid dynamical systems [4]. Due to their ability to exhibit complex dynamical behavior, hybrid NSDDEs are widely applied in various fields, such as being used to simulate the signal transmission and inter-neuron interactions of neural networks in biomedicine, and for designing control systems to achieve precise control of complex systems in engineering. In recent years, the stability of hybrid NSDDEs has received much attention. There have been a number of achievements on the issue of hybrid NSDDEs [5].
In real-world scenarios, many dynamical systems are usually subjected to random abrupt changes caused by different kinds of environmental noise [6]. Typically, dynamical systems with white noise perturbations are modeled by Itˆo stochastic differential equations (SDEs) [7]. Research on the stability analysis of SDEs has been abundant up to now [8,9]. However, sensor noise in engineering applications is usually time-correlated, and white noise models cannot accurately capture this characteristic. Moreover, the noise intensity is often related to the system state, such as the power of thermal noise in circuits varying with temperature. Therefore, introducing colored noise can better describe the spectral characteristics of real-world noise. As a result, dynamical systems with colored noise are typically described using SDEs where the noise has finite second-order moments. Such models can better capture the non-linearities and correlations that exist in many natural systems, thus enhancing the understanding and explanation of the behavior for various systems. Lately, the noise-to-state stability (NSS) of hybrid SDEs with colored noise was studied in [10], and the NSS of stochastic impulse-delayed systems with multiple random impulses was discussed in [11].
Currently, The majority of stability criteria apply only to stochastic systems where the coefficients meet the linear growth condition (LGC). Currently, The majority of stability criteria apply only to stochastic systems where the coefficients meet the linear growth condition (LGC). However, the nonlinear dynamic behaviors in real-world systems [12], such as the Duffing equation in mechanical vibrations or the nonlinear rate equations in chemical reaction networks, require model coefficients to satisfy polynomial growth conditions (PGC) [13], rather than the traditional LGC. As research has advanced, researchers have increasingly focused on the stability of highly nonlinear SDEs as research has progressed [14,15]. For example, the stability of hybrid variable multiple-delay SDEs, which are highly nonlinear, was considered in [16], and the stability of hybrid NSDDEs under PGC has been addressed in [17]. As we all know, these stability criteria can generally be divided into delay-independent stability (DIS) and delay-dependent stability (DDS) [18]. The DDS criterion contains information about time delay, considering the size of time delay, and is therefore generally less conservative than the DIS criterion, which is suitable for time delay of any size [19]. There are many theoretical results about DDS for SDEs [20,21]. Recently, the DDS of highly nonlinear hybrid NSDDEs was studied in [22], while the DDS criterion for hybrid NSDDEs was derived using Lyapunov functionals in [23].
In fields such as engineering control, biological neural networks, and environmental science, neutral stochastic differential systems are often subject to the coupled influence of multiple factors, including time delay effect, colored noise, high nonlinearity, and Markov switching mechanisms. This complexity imposes higher demands on model construction. However, existing models are largely constrained by linear growth conditions and white noise assumptions, neglecting the dynamic interplay among time delay, noise, and switching behaviors, which results in insufficient accuracy in modeling real systems. Therefore, there is an urgent practical need to develop neutral stochastic differential delay models that integrate high nonlinearity, colored noise, and Markov switching.
To better explain our purpose, consider the voltage regulation problem in power systems, where the dynamical behavior is affected by equipment failures (Markov switching) and environmental vibrations (colored noise). The system dynamics can be modeled as hybrid NSDDEs with colored noise, as follows:
d[Φ(t)−W(Φ(t−δ))]=f(Φ(t),Φ(t−δ),π(t),t)dt+σ(Φ(t),Φ(t−δ),π(t),t)ξ(t)dt, | (1.1) |
where W(Φ(t−δ))=0.1Φ(t−δ), π(t) is a Markov chain taking values from the set S = {1,2}, with π(t)=1 representing the normal mode and π(t)=2 representing the failure mode, and its generator matrix given by Γ=[−3,3;1,−1]. We generate ξ(t)∈R using the formula ξ(t)=0.5cos(2t+ϖ), where ϖ is a uniformly distributed random variable in the interval [0,2π] and Eξ(t)2≤0.125. We define
f(Φ,ν,i,t)={−6Φ3−1.5ν,i=1,−6Φ3−1ν,i=2,σ(Φ,ν,i,t)={0.1ν,i=1,0.2ν,i=2. |
If δ takes a value of 0.015, it can be observed from Figure 1 that the highly nonlinear hybrid NSDDEs (1.1) are asymptotically stable. In contrast, if δ is set to 2, Figure 2 shows that the same highly nonlinear hybrid NSDDEs (1.1) become unstable. Put differently, the size of time delay affects the stability of system (1.1). However, for the highly nonlinear hybrid NSDDEs with colored noise, there are few DDS criteria that can be utilized to obtain a sufficient bound on the time delay δ and ensure the stability of its solution. Therefore, the focus of this paper is on exploring a class of highly nonlinear hybrid NSDDEs with colored noise and establishing applicable DDS criteria.
The primary contributions are summarized as follows: (1) Colored noise is introduced in hybrid NSDDEs, and the coefficients of hybrid NSDDEs are highly nonlinear. (2) The existence of a global solution for highly nonlinear hybrid NSDDEs with colored noise is proved under PGC. (3) The Lyapunov functional considered in this paper involves time delay, which makes our stability criteria delay-dependent and thus less conservative.
Notations: If Φ∈Rn, |Φ| represents its Euclidean norm. The set of continuous functions ϱ:[−δ,0]→Rn is denoted by C([−δ,0];Rn) for δ>0, with its norm defined as ‖ϱ‖=sup−δ≤u≤0|ϱ(u)|. Let C1,1(Rn×S×R+;R+) represent the family of all continuous functions U(Φ,i,t) that are continuously differentiable once with respect to Φ and t, respectively. The family of all quasi-polynomial functions H(ι) with non-negative continuous coefficients are defined as H(Rn;R+), and H(ι) is expressed as H(ι)=ak|ι|dk+ak−1|ι|dk−1+⋯+a1|ι|d1 with ai≥0 (i=1,2,⋯,k) and dk≥dk−1≥⋯≥d1≥1. A continuous function β∈C(R+;R+) is considered to belong to the set of K-function if it is strictly increasing and β(0)=0. If β(⋅) is also radially unbounded, then it is said to belong to the set of KR-functions. Additionally, a function Ξ(Φ,t)∈C(R+×R+;R+) is considered to belong to the set of KL-functions if it is a K-function for every fixed t and decreases to zero for every fixed Φ as t→∞.
Suppose (Ω,F,{Ft}t≥0,P) is a complete probability space, where {Ft}t≥0 is a filtration that satisfies right continuity, and F0 contains all P-null sets. For any t≥0, let π(t) be a right-continuous Markov chain on the complete probability space with state space S={1,2,…,N} and generator Γ=[γij]N×N. Here, γij≥0 and γii=−∑Nj=1,j≠iγij≤0.
Next, we analyze the given highly nonlinear hybrid NSDDE with colored noise
d[Φ(t)−W(Φ(t−δ))]=f(Φ(t),Φ(t−δ),π(t),t)dt+σ(Φ(t),Φ(t−δ),π(t),t)ξ(t)dt, | (2.1) |
and initial condition
{Φ(t):−δ≤t≤0}=η∈LαF0([−δ,0];Rn),π(0)=π0∈S, | (2.2) |
where Φ(t)∈Rn denotes the state vector, and ξ(t)∈Rd represents colored noise. The f∈C(Rn×Rn×S×R+;Rn), σ∈C(Rn×Rn×S ×R+;Rn×d) and W∈C(Rn;Rn) denote Borel-measurable functions.
In the following, we provide some assumptions for (2.1).
Assumption 2.1. [17]. For any h>0 and for all ˜u,˜v,ˉu,ˉv∈Rn, where |˜u|∨|˜u|∨|ˉu|∨|ˉv|≤h, there exists a constant Lh>0 such that
|f(˜u,˜v,i,t)−f(ˉu,ˉv,i,t)|∨|σ(˜u,˜v,i,t)−σ(ˉu,ˉv,i,t)|≤Lh(|˜u−ˉu|+|˜v−ˉv|) | (2.3) |
with (i,t)∈S×R+.
Assumption 2.2. [17]. For any ˜u,˜v∈Rn, there are constants Q>0, α1>1 and α2≥1 satisfying
|f(˜u,˜v,i,t)|≤Q(1+|˜u|α1+|˜v|α1),|σ(˜u,˜v,i,t)|≤Q(1+|˜u|α2+|˜v|α2) | (2.4) |
with (i,t)∈S×R+. Furthermore, there also is a constant ˜ω∈(0,√22) satisfying
|W(˜u)−W(˜v)|≤˜ω|˜u−˜v| | (2.5) |
with W(0)=0.
Remark 2.1. Assumptions 2.1 and 2.2 ensure that the coefficients f and σ satisfy the local Lipschitz condition and the PGC.
Remark 2.2. Assumption 2.2 in condition (2.5) shows that the function W is globally Lipschitz continuous and satisfies the LGC: |W(˜u)|≤˜ω|˜u|.
Assumption 2.3 [10]. Given the process ξ(t) is both piecewise continuous and Ft-adapted. Furthermore, it satisfies sup0≤s≤tE|ξ(t)|2<∞.
Remark 2.3. By Assumption 2.3, for any t≥0, it can be checked that ξ(t)<∞ almost surely (a.s.).
For convenience, we assume that α1>1, although it is sufficient to have only max{α1,α2}>1. The PGC (2.4) is referred to as Assumption 2.2, and it is well-known that under Assumptions 2.1–2.3, the hybrid NSDDE (2.1) has a unique maximal local solution, but this solution may blow up in finite time. To prevent this phenomenon, some restrictions are given below.
Assumption 2.4. Let ˉU∈C1,1(Rn×S×R+;R+) and H∈H(Rn,R+). γ(⋅)∈KR and is convex, along with b1,b2,b3>0 and α≥2(α1∨α2), such that
b3<b2,|Φ|α≤ˉU(Φ,i,t)≤H(Φ), | (2.6) |
and
dˉU(Φ−W(Φ(t−δ)),i,t)=ˉUt(Φ−W(Φ(t−δ)),i,t)+ˉUΦ(Φ−W(Φ(t−δ)),i,t)f(Φ,ν,i,t)+ˉUΦ(Φ−W(Φ(t−δ)),i,t)σ(Φ,ν,i,t)ξ(t)+N∑j=1γijˉU(x−W(Φ(t−δ)),j,t)≤b1−b2H(Φ)+b3H(ν)+γ(|ξ(t)|2) | (2.7) |
for any (Φ,ν,i,t)∈Rn×Rn×S×R+.
Remark 2.4. Assumption 2.4 is the key to the presence of a global solution for hybrid NSDDE (2.1) in the nonlinear scenario.
Remark 2.5. Assumption 2.4 is an improvement of Assumption 2.4 in [10] since this paper assumes that W(⋅) satisfies Assumption 2.1. Therefore, Assumption 2.4 is valid in this paper.
Definition 2.1. [24]. For α>0, assume that Ξ∈KL and β∈K exist, satisfying
E|x(t)|α≤Ξ(||η||,t)+β(sup0≤s≤tE|ξ(s)|2) |
where t∈R+ and η∈LαF0([−δ,0];Rn). Then hybrid NSDDE (2.1) is said to be NSS in the αth moment (NSS-α-M). In particular, when α=2, it is commonly referred to as NSS in the mean square.
Lemma 2.1. [25]. If Assumption 2.2 is satisfied and there exists a constant α≥1, then
|˜u−W(˜v)|α≤(1+˜ω)α−1(|˜u|α+˜ω|˜v|α),|˜u|α≤˜ω|˜v|α+|˜u−W(˜v)|α(1−˜ω)α−1. |
holds, where ˜u,˜v∈Rn.
This section presents a sufficient condition for proving the existence of a unique global solution to hybrid NSDDE (2.1). Additionally, it explores the NSS and DDS criteria for global solutions.
Theorem 3.1. Assuming that Assumptions 2.1–2.4 are satisfied, we can make the following assertions for hybrid NSDDE (2.1).
(ⅰ) Hybrid NSDDE (2.1) has a unique global solution on the interval [−δ,∞).
(ⅱ) The global solution satisfies
lim supt→∞E|Φ(t)|α≤1λ(1−√˜ω)(1−˜ω)α−1[b1+γ(sup0≤s≤tE|ξ(s)|2)], | (3.1) |
∀t∈R+, where η∈LαF0([−δ,0];Rn) and ˉλ>0 is the only solution of
b2−λ2dk−1−eλδ(b3+λ2dk−1)=0, | (3.2) |
where dk=deg(H(x)).
(ⅲ) When b1=0 and η∈LαF0([−δ,0];Rn), the global solution satisfies
E|Φ(s)|α≤M0e−λt+1λ(1−√˜ω)(1−˜ω)α−1γ(sup0≤s≤tE|ξ(s)|2) | (3.3) |
where t∈R+, M0=√˜ω1−√˜ωE||η||α+1(1−√˜ω)(1−˜ω)α−1CλH(||η||). In other words, the global solution of hybrid NSDDE (2.1) is NSS-α-M.
Proof. To better understand the proof process, we can illustrate it in three steps.
Step 1. By relying on Assumptions 2.1–2.3, it can be easily demonstrated that hybrid NSDDEs (2.1) possesses a unique maximal local solution on the interval [−δ,φ∞), where φ∞ represents the explosion time. We choose an integer ˉh0>0 that is large enough to ensure ||η||≤ˉh0. We define the stopping time ϕˉh=inf{t∈[0,φ∞):|Φ(t)|≥ˉh} for every integer ˉh≥ˉh0, where inf∅=∞. It is an obvious fact that ϕˉh increases as ˉh→∞ and ϕ∞=limˉh→∞ϕˉh≤φ∞ a.s. If ϕ∞=∞ a.s., in that case, there is one unique global solution for hybrid NSDDE (2.1) on the interval [−δ,φ∞).
We can obtain from (2.6) and (2.7) that
EˉU(Φ(t∧ϕˉh)−W(Φ(t∧ϕˉh−δ)),π(t∧ϕˉh),t∧ϕˉh)≤H(Φ(0)−W(Φ(−δ)))+b1t−b2E∫t∧ϕˉh0H(Φ(s))ds+b3E∫t∧ϕˉh0H(Φ(s−δ))ds+E∫t∧ϕˉh0γ(|ξ(s)|2)ds. | (3.4) |
Based on the information about the time delay, one gets
∫t∧ϕˉh0H(Φ(s−δ))ds≤∫0−δH(Φ(s))ds+∫t∧ϕˉh0H(Φ(s))ds. | (3.5) |
Substituting (3.5) into (3.4) and applying the Jensen inequality, one has
EˉU(Φ(t∧ϕˉh)−W(Φ(t∧ϕˉh−δ)),π(t∧ϕˉh),t∧ϕˉh)≤M1+(b1+γ(sup0≤s≤tE|ξ(s)|2))t−(b2−b3)E∫t∧ϕˉh0H(Φ(s))ds | (3.6) |
with M1=H((1+˜ω)||η||)+b3δH(||η||). Combining (2.6) and (3.6), we can deduce
E|Φ(t∧ϕˉh)−W(Φ(t∧ϕˉh−δ))|α≤M1+(b1+γ(sup0≤s≤tE|ξ(s)|2))t. |
Let us define μˉh=inf|y|≥(1−˜ω)ˉh,t≥0|y|α. In accordance with the definition of ϕˉh, for t∈[−δ,ϕˉh], one has |Φ(t)|≤ˉh. We observe that
|Φ(ϕˉh)−W(Φ(ϕˉh−δ))|I{ϕˉh≤t}≥(|Φ(ϕˉh)|−|W(Φ(ϕˉh−δ))|)I{ϕˉh≤t}≥(|Φ(ϕˉh)|−˜ω|Φ(ϕˉh−δ)|)I{ϕˉh≤t}≥ˉh−˜ωˉh=(1−˜ω)ˉh. |
Noting that
E|Φ(ϕˉh∧t)−W(Φ(ϕˉh∧t−δ))|α≥E[|Φ(ϕˉh)−W(Φ(ϕˉh−δ))|αI{ϕˉh≤t}]≥E[inf|y|≥(1−˜ω)ˉh,t≥0|y|αI{ϕˉh≤t}]=μˉhP{ϕˉh≤t}, |
we see that
μˉhP{ϕˉh≤t}≤M1+(b1+γ(sup0≤s≤tE|ξ(s)|2))t. |
Clearly, one obtains limˉh→∞μˉh=∞. Letting ˉh→∞, we have P{ϕ∞≤t}=0, which in turn leads to P{ϕ∞>t}=1. As we let t→∞, we find that P{ϕ∞=∞}=1, which means that ϕ∞=∞ a.s. Therefore, we can conclude that assertion (i) holds as required.
Step 2. Since H∈H(Rn;R+), we set H(ι)=ak|ι|dk+ak−1|ι|dk−1+⋯+a1|ι|d1. Combining Lemma 2.1 and ˜ω∈(0,√22), we derive
H(Φ(s)−W(Φ(s−δ)))=ak|Φ(s)−W(Φ(s−δ))|dk+ak−1|Φ(s)−W(Φ(s−δ))|dk−1+⋯+a1|Φ(s)−W(Φ(s−δ))|d1≤ak(1+˜ω)dk−1(|Φ(s)|dk+˜ω|Φ(s−δ)|dk)+⋯+a1(1+˜ω)d1−1(|Φ(s)|d1+˜ω|Φ(s−δ)|d1)≤2dk−1[ak(|Φ(s)|dk+|Φ(s−δ)|dk)+⋯+a1(|Φ(s)|d1+|Φ(s−δ)|d1)]≤2dk−1(H(Φ(s))+H(Φ(s−δ))). |
By using the zero-point theorem and (2.6), it can be concluded that Eq (3.2) has a unique solution ˉλ>0. For any λ∈(0,ˉλ∧12δlog(1˜ω)], we get
Eeλ(t∧ϕˉh)ˉU(Φ(t∧ϕˉh)−W(Φ(t∧ϕˉh−δ)),π(t∧ϕˉh),t∧ϕˉh)≤ˉU(Φ(0)+W(Φ(−δ)),π(0),0)+E∫t∧ϕˉh0eλs[b1+λH(Φ(s)−W(Φ(s−δ))) |
−b2H(Φ(s))+b3H(Φ(s−δ))]ds+E∫t∧ϕˉh0eλsγ(|ξ(s)|2)ds≤E∫t∧ϕˉh0eλs(2dk−1λ(H(Φ(s))+H(Φ(s−δ)))+b1−b2H(Φ(s))+b3H(Φ(s−δ)))ds+E∫t∧ϕˉh0eλsγ(|ξ(s)|2)ds≤E∫t∧ϕˉh0eλs(b1+(λ2dk−1−b2)H(Φ(s))+(b3+λ2dk−1)H(Φ(s−δ)))ds+E∫t∧ϕˉh0eλsγ(|ξ(s)|2)ds. | (3.7) |
Since
∫t∧ϕˉh0eλsH(Φ(s−δ))ds=eλδ∫0−δeλsH(Φ(s))ds+eλδ∫t∧ϕˉh0eλsH(Φ(s))ds, |
from (3.2) and (3.7), one has
Eeλ(t∧ϕˉh)ˉU(Φ(t∧ϕˉh)−W(Φ(t∧ϕˉh−δ)),π(t∧ϕˉh),t∧ϕˉh)−ˉU(Φ(0)−W(Φ(−δ)),π(0),0)≤1λ[b1+γ(sup0≤s≤tE|ξ(s)|2)]eλt+eλδ(b3+λ2dk−1)E∫0−δeλsH(Φ(s))ds. | (3.8) |
From the Fatou lemma and (2.6), it follows that (3.8) yields
Eeλt|Φ(t)−W(Φ(t−δ))|α≤CλH(||η||)+1λ[b1+γ(sup0≤s≤tE|ξ(s)|2)]eλt, | (3.9) |
where Cλ=[2dk+eλδδ(b3+λ2dk−1)]. It is also evident from (3.9) and Lemma 2.1 that
eλtE|Φ(t)|α≤˜ωeλtE|Φ(t−δ)|α+1(1−˜ω)α−1eλtE|Φ(t)−W(Φ(t−δ))|α≤√˜ωeλ(t−δ)E|Φ(t−δ)|α+1(1−˜ω)α−1{CλH(||η||)+1λ[b1+γ(sup0≤s≤tE|ξ(s)|2)]eλt}. |
Thus, we get
sup0≤s≤teλsE|Φ(s)|α≤√˜ω(E||η||α+sup0≤s≤teλsE|Φ(s)|α) |
+1(1−˜ω)α−1{CλH(||η||)+1λ[b1+γ(sup0≤s≤tE|ξ(s)|2)]eλt}. |
Thus,
sup0≤s≤teλsE|Φ(s)|α≤√˜ω1−√˜ωE||η||α+1(1−√˜ω)(1−˜ω)α−1CλH(||η||)+1λ(1−√˜ω)(1−˜ω)α−1[b1+γ(sup0≤s≤tE|ξ(s)|2)]eλt. |
In particular,
E|Φ(t)|α≤M0e−λt+1λ(1−√˜ω)(1−˜ω)α−1⋅[b1+γ(sup0≤s≤tE|ξ(s)|2)], | (3.10) |
where M0=√˜ω1−√˜ωE||η||α+1(1−√˜ω)(1−˜ω)α−1CλH(||η||). Hence, setting t→∞ yields the following inequality:
lim supt→∞E|Φ(t)|α≤1λ(1−√˜ω)(1−˜ω)α−1[b1+γ(sup0≤s≤tE|ξ(s)|2)], |
showing that assertion (ⅱ) is satisfied.
Step 3. When b1=0, (3.10) still holds. Thus, when b1=0, there holds
E|Φ(t)|α≤M0e−λt+1λ(1−√˜ω)(1−˜ω)α−1γ(sup0≤s≤tE|ξ(s)|2). |
By Definition 2.1, we can easily know that the global solution of hybrid NSDDE (2.1) is NSS-α-M. As a result, we can infer that the expected assertion (ⅲ) is valid.
Assumption 3.1. Given that ζ(t) is both piecewise continuous and Ft-adapted, one can conclude the existence of a positive scalar μ such that supt≥0E|ξ(t)|2<μ.
Remark 3.1. To discuss the asymptotic properties of the global solution for hybrid NSDDE (2.1), a stricter assumption about the colored noise ξ(t), namely Assumption 3.1, is required. It is evident that when Assumption 3.1 holds, Assumption 2.3 also holds. Therefore, under the conditions that Assumptions 2.1, 2.2, 2.4, and 3.1 are satisfied, the conclusions in Theorem 3.1 still hold for hybrid NSDDE (2.1).
Next, for t∈R+, we define ˉΦt={Φ(t+ζ): −2δ≤ζ≤0} and ˉπt={π(t+ζ):−2δ≤ζ≤0}. Furthermore, let Φ(ζ)=η(−δ) for ζ∈[−2δ,−δ) and π(ζ)=π0 for ζ∈[−2δ,0). For all Φ,ν∈Rn and (i,t)∈S×[−2δ,0), let f(Φ,ν,i,ζ)=f(Φ,ν,i,0) as well as σ(Φ,ν,i,ζ)=σ(Φ,ν,i,0). Define the following delay-dependent Lyapunov functional:
V(ˉΦt,ˉπt,t)=U(Φ(t)−W(Φ(t−δ)),π(t),t)+θ∫0−δ∫tt+sF(u)duds, |
where U∈C1,1(Rn×S×R+;R+) satisfies lim|Φ|→∞[inf(t,i)∈R+×SU(Φ,i,t)]=∞, θ>0 is a constant that requires identification, and F(u)=δ|f(Φ(u),Φ(u−δ),π(u),u)|2+μδ|σ(Φ(u),Φ(u−δ),π(u),u)|2. Then, we have
dV(ˉΦt,ˉπt,t)=UΦ(Φ(t)−W(Φ(t−δ),π(t),t)×[f(Φ(t),Φ(t−δ),π(t),t)−f(Φ(t),Φ(t),π(t),t)]+LU(Φ(t)−W(Φ(t−δ),Φ(t−δ),π(t),t))+θδF(t)−θ∫tt−δF(u)du, | (3.11) |
where
LU(Φ−W(Φ(t−δ)),Φ(t−δ),i,t)=Ut(Φ−W(Φ(t−δ)),i,t)+UΦ(Φ−W(Φ(t−δ)),i,t)[f(Φ,Φ,i,t)+σ(Φ,Φ(t−δ),i,t)ξ(t)]+N∑j=1γijU(Φ−W(Φ(t−δ)),j,t). | (3.12) |
In order to analyze the DDS of hybrid NSDDE (2.1), additional assumptions are required.
Assumption 3.2. Consider the functions U∈C1,1(Rn×S×R+;R+), U1∈H(Rn;R+), G∈C(Rn;R+), and the constants βi>0 (i=1,2,3) and ϑk>0 (k=1,2) satisfying
β1>β2+μβ3 | (3.13) |
and
LU(Φ−W(ν),ν,i,t)+ϑ1|UΦ(Φ−W(ν),i,t)|2+ϑ2|f(Φ,ν,i,t)|2+μϑ2|σ(Φ,ν,i,t)|2≤−β1U1(Φ)+β2U1(ν)+β3U1(Φ)|ξ(t)|2−G(Φ−W(ν)) | (3.14) |
for all (Φ,ν,i,t)∈Rn×Rn×S×R+. In addition, G also satisfies the following condition:
G(Φ)=0only whenΦ=0. | (3.15) |
Assumption 3.3. Assume there exists a constant L>0 satisfying the following inequality:
|f(Φ,Φ,i,t)−f(Φ,ˉΦ,i,t)|≤L|Φ−ˉΦ| | (3.16) |
where (Φ,ˉΦ,i,t)∈Rn×Rn×S×R+.
Remark 3.2. Assumption 3.2 imposes the necessary requirement on the operator L. Assumption 3.3 states that f satisfies the Lipschitz condition.
Theorem 3.2. Under Assumptions 2.1, 2.2, 2.4, and 3.1–3.3, the condition
L2δ2≤(1−2˜ω2)ϑ1ϑ2 | (3.17) |
holds, which implies that the solution to the hybrid NSDDE (2.1) satisfies the following conditions:
∫∞0EU1(Φ(t))dt<∞, | (3.18) |
sup0≤t<∞EU(Φ(t)−W(Φ(t−δ)),π(t),t)<∞. | (3.19) |
Proof. Let ρh=inf{t≥0:|Φ(t)−W(Φ(t−δ))|≥h}. Using the ordinary differential formula, we obtain
EV(ˉΦt∧ρh,ˉπt∧ρh,t∧ρh)=V(ˉΦ0,ˉπ0,0)+E∫t∧ρh0dV(ˉΦs,ˉπs,s). | (3.20) |
Let θ=L2/(ϑ1(1−2˜ω2)). From Assumption 3.3, there holds
UΦ(Φ(t)−W(Φ(t−δ),π(t),t)×[f(Φ(t),Φ(t−δ),π(t),t)−f(Φ(t),Φ(t),π(t),t)]≤ϑ1|UΦ(Φ(t)−W(Φ(t−δ),π(t),t)|2+L24ϑ1|Φ(t)−Φ(t−δ)|2. | (3.21) |
According to condition (3.17), it is not difficult to get θδ2≤ϑ2. Then, combining (3.11), (3.14), and (3.21), we have
dV(ˉΦs,ˉπs,s)≤LU(Φ(s)−W(Φ(s−δ)),Φ(s−δ),π(s),s)+ϑ1|UΦ(Φ(t)−W(Φ(t−δ),π(t),t)|2+L24ϑ1|Φ(t)−Φ(t−δ)|2+ϑ2|f(Φ(s),Φ(s−δ),π(s),s)|2+μϑ2|σ(Φ(s),Φ(s−δ),π(s),s)|2−L2ϑ1(1−2˜ω2)∫ss−δF(u)du≤−β1U1(Φ(s))+β2U1(Φ(s−δ))+β3U1(Φ(s))|ξ(s)|2−G(Φ(s)−W(Φ(s−δ)))+L24ϑ1|Φ(t)−Φ(t−δ)|2−L2ϑ1(1−2˜ω2)∫ss−δF(u)du. |
Substituting this into (3.20) gives
EV(ˉΦt∧ρh,ˉπt∧ρh,t∧ρh)≤V(ˉΦ0,ˉπ0,0)+C1−C2+C3−C4, | (3.22) |
where
C1=E∫t∧ρh0[−β1U1(Φ(s))+β2U1(Φ(s−δ))+β3U1(Φ(s))|ξ(s)|2]ds,C2=E∫t∧ρh0G(Φ(s)−W(Φ(s−δ)))ds,C3=L24ϑ1E∫t∧ρh0|Φ(s)−Φ(s−δ)|2ds,C4=L2ϑ1(1−2˜ω2)E∫t∧ρh0∫ss−δF(u)duds. |
Noting that
∫t∧ρh0U1(Φ(s−δ))ds≤∫t∧ρh−δU1(Φ(u))du≤∫0−δU1(Φ(u))du+∫t∧ρh0U1(Φ(u))du, | (3.23) |
it yields from (3.23) that
C1≤β2∫0−δU1(Φ(s))ds−(β1−β2−μβ3)E∫t∧ρh0U1(Φ(s))ds. |
Bringing this into (3.22) leads to
E∫t∧ρh0U1(Φ(s))ds≤1β1−β2−μβ3[V(ˉΦ0,ˉπ0,0)+β2E∫0−δU1(Φ(s))ds−C2+C3−C4]. | (3.24) |
As we let h→∞ and apply the Fatou lemma along with the Fubini theorem to (3.24), we derive
E∫t0U1(Φ(s))ds≤1β1−β2−μβ3×[K1−ˉC2+ˉC3−ˉC4], | (3.25) |
where
K1=V(ˉΦ0,ˉπ0,0)+β2E∫0−δU1(Φ(s))ds,ˉC2=E∫t0G(Φ(s)−W(Φ(s−δ)))ds,ˉC3=L24ϑ1∫t0E|Φ(s)−Φ(s−δ)|2ds,ˉC4=L2ϑ1(1−2˜ω2)E∫t0∫ss−δF(u)duds. |
Considering that G∈C(Rn;R+), we can deduce from (3.25) that
E∫t0U1(Φ(s))ds≤1β1−β2−μβ3×[K1+ˉC3−ˉC4]. | (3.26) |
On the one hand, for t∈[0,δ], one has
ˉC3≤L22ϑ1∫δ0(E|Φ(s)|2+E|Φ(s−δ)|2)ds≤δL2ϑ1(sup−δ≤v≤δE|Φ(v)|2)=:K2. |
On the other hand, for t>δ, we get
ˉC3≤K2+L24ϑ1∫tδE|Φ(s)−Φ(s−δ)|2ds. |
Combining (2.1) and (2.5) results in
|Φ(s)−Φ(s−δ)|≤|[Φ(s)−W(Φ(s−δ))]−[Φ(s−δ)−W(Φ(s−2δ))]|+|W(Φ(s−δ))−W(Φ(s−2δ))|≤˜ω|Φ(s−δ)−Φ(s−2δ)|+|∫ss−δf(Φ(u),Φ(u−δ),π(u),u)+σ(Φ(u),Φ(u−δ),π(u),u)ξ(u)du|. |
Hence, together with Assumption 3.1, we obtain
E|Φ(s)−Φ(s−δ)|2≤2˜ω2E|Φ(s−δ)−Φ(s−2δ)|2+2E|∫ss−δf(Φ(u),Φ(u−δ),π(u),u)+σ(Φ(u),Φ(u−δ),π(u),u)ξ(u)du|2≤2˜ω2E|Φ(s−δ)−Φ(s−2δ)|2+4E∫ss−δF(u)du, |
which implies
\begin{aligned} \int_{\delta}^{t}\mathbb{E}\vert \Phi(s)- \Phi(s-\delta)\vert^2ds\leq& 2 \tilde{\omega }^2\int_{\delta}^{t}\mathbb{E}\vert \Phi(s-\delta)- \Phi(s-2\delta)\vert^2ds\\ &+4\mathbb{E}\int_{\delta}^{t}\int_{s-\delta}^{s}F(u)duds\\ \leq&2 \tilde{\omega }^2\int_{0}^{t}\mathbb{E}\vert \Phi(s)- \Phi(s-\delta)\vert^2ds\\ &+4\mathbb{E}\int_{\delta}^{t}\int_{s-\delta}^{s}F(u)duds. \end{aligned} |
Noting that 0 < \kappa < \frac{\sqrt{2}}{2} , then
\begin{aligned} \int_{\delta}^{t}\mathbb{E}\vert \Phi(s)- \Phi(s-\delta)\vert^2ds&\leq \frac{2 \tilde{\omega }^2}{1-2 \tilde{\omega }^2}\int_{0}^{\delta}\mathbb{E}\vert \Phi(s)- \Phi(s-\delta)\vert^2ds\\ &\quad+\frac{4}{1-2 \tilde{\omega }^2}\mathbb{E}\int_{\delta}^{t}\int_{s-\delta}^{s}F(u)duds. \end{aligned} |
Hence,
\begin{align} \begin{aligned} \bar{C}_3\leq& K_2+\frac{L^2}{4\vartheta_{1}}\bigg(\frac{2 \tilde{\omega }^2}{1-2 \tilde{\omega }^2}\int_{0}^{\delta}\mathbb{E}\vert \Phi(s)- \Phi(s-\delta)\vert^2ds+\frac{4}{1-2 \tilde{\omega }^2}\mathbb{E}\int_{\delta}^{t}\int_{s-\delta}^{s}F(u)duds\bigg)\\ \leq& K_2+\frac{2 \tilde{\omega }^2\delta L^2}{(1-2 \tilde{\omega }^2)\vartheta_{1}}\sup\limits_{-\delta\leq v\leq\delta}\mathbb{E}\vert \Phi(v)\vert^2+\bar{C}_4\\ = &K_3+\bar{C}_4, \end{aligned} \end{align} | (3.27) |
where K_3 = K_2+\frac{2 \tilde{\omega }^2\delta L^2}{(1-2 \tilde{\omega }^2)\vartheta_{1}}\sup_{-\delta\leq v\leq\delta}\mathbb{E}\vert \Phi(v)\vert^2 . Bringing (3.27) into (3.26) and letting t\to\infty , we derive
\begin{align} \begin{aligned} \mathbb{E}\int_{0}^{\infty}U_1( \Phi(s))ds\leq\frac{1}{\beta_{1}-\beta_{2}-\mu\beta_{3}}(K_1+K_3). \end{aligned} \end{align} | (3.28) |
Applying the Fubini theorem again to (3.27) yields the result (3.18). Letting h\to\infty and combining (3.20), (3.22), and (3.27), we calculate
\mathbb{E}U( \Phi(t)- W( \Phi(t-\delta)), \pi(t),t)\leq K_1+K_3 < \infty, |
which indicates
\sup\limits_{0\leq t < \infty}\mathbb{E}U( \Phi(t)- W( \Phi(t-\delta)), \pi(t),t) < \infty. |
Hence, (3.19) holds.
Corollary 3.1. Suppose that the conditions of Theorem 3.2 are true and that there exist two constants d > 0 and \hat{\alpha} > 0 , satisfying
\begin{aligned} d\vert \Phi\vert^{\hat{\alpha}}\leq U_1( \Phi) \end{aligned} |
for any \Phi\in \mathcal{R}^n . Then, we can obtain the solution of the hybrid NSDDE (2.1), satisfying
\begin{align} \begin{aligned} \int_{0}^{\infty}\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}dt < \infty. \end{aligned} \end{align} | (3.29) |
Namely, hybrid NSDDE (2.1) is H_{\infty} -stable in L^{\hat{\alpha}} .
Remark 3.3. Theorem 3.1 proves that NSDDE (2.1) possesses NSS- \alpha -M. This result describes the asymptotic behavior of system states under the influence of noise and tends to be stable under certain conditions. Theorem 3.2 further establishes the integral boundedness of the function U_1(\Phi) , that is,
\int_{0}^{\infty}\mathbb{E}U_{1}( \Phi(t))dt < \infty, |
which demonstrates that the cumulative energy of the system state over time is finite. Corollary 3.1 states that NSDDE (2.1) is H_{\infty} -stable in L^{\hat{\alpha}} . This is a special case of Theorem 3.2. Specifically, when d\vert \Phi\vert^{\hat{\alpha}}\leq U_1(\Phi) , the integral boundedness of U_{1}(\Phi(t)) directly implies the integral boundedness of \vert \Phi\vert^{\hat{\alpha}} , that is,
\begin{aligned} \int_{0}^{\infty}\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}dt < \infty, \end{aligned} |
thereby ensuring that NSDDE (2.1) is H_{\infty} -stable in L^{\hat{\alpha}} .
Next, we establish a theorem regarding the asymptotic stability in L^{\hat{\alpha}} for hybrid NSDDE (2.1).
Theorem 3.3. Suppose that the conditions of Corollary 3.1 are true. If {\hat{\alpha}}\geq2 and 2({\hat{\alpha}}-1)\vee({\hat{\alpha}}+\alpha_1-1)\vee2({\hat{\alpha}}+\alpha_2-1)\leq \alpha , then the solution of hybrid NSDDE (2.1) satisfies
\lim\limits_{t\to\infty}\mathbb{E}\vert \Phi\vert^{\hat{\alpha}} = 0. |
Namely, hybrid NSDDE (2.1) is asymptotically stability in L^{\hat{\alpha}} .
Proof. Using this inequality \vert \bar{a}\bar{b}\vert\leq \bar{c}\vert \bar{a}\vert^2+\frac{1}{4\bar{c}}\vert \bar{b}\vert^2 with any \bar{a}, \bar{b}\in \mathcal{R} \ \text{and}\ \bar{c} > 0 . For any 0\leq t_1 < t_2 < \infty , from Assumptions 2.2 and 3.1, there holds
\begin{aligned} &\vert \mathbb{E} \vert \Phi(t_2)- W( \Phi(t_2-\delta))\vert ^{\hat{\alpha}}-\mathbb{E}\vert \Phi(t_1)- W( \Phi(t_1-\delta ))\vert ^{\hat{\alpha}}\vert\\ = &\vert\mathbb{E}\int_{t_1}^{t_2}{\hat{\alpha}}\vert \Phi(t)- W( \Phi(t-\delta ))\vert^{{\hat{\alpha}}-1}\bigg(f( \Phi(t), \Phi(t-\delta), \pi(t),t)\\ &+ \sigma( \Phi(t), \Phi(t-\delta ), \pi(t),t) \xi(t)\bigg)dt\vert\\ \leq&\mathbb{E}\int_{t_1}^{t_2}\bigg({\hat{\alpha}}Q\vert \Phi(t)- W( \Phi(t-\delta))\vert ^{{\hat{\alpha}}-1}\times\left(1+\vert \Phi\vert^{\alpha_1}+\vert \Phi(t-\delta)\vert^{\alpha_1}\right)\\ &+{\hat{\alpha}}Q\vert \Phi(t)- W( \Phi(t-\delta))\vert^{{\hat{\alpha}}-1}\times\left(1+\vert \Phi\vert^{\alpha_2}+\vert \Phi(t-\delta)\vert^{\alpha_2}\right)\vert \xi(t)\vert \bigg)dt\\ \leq& \mathbb{E}\int_{t_1}^{t_2}\bigg({\hat{\alpha}}Q\vert \Phi(t)- W( \Phi(t-\delta ))\vert ^{{\hat{\alpha}}-1}\times\left(1+\vert \Phi\vert^{\alpha_1}+\vert \Phi(t-\delta)\vert^{\alpha_1}\right)\\ &+\bar{c}{\hat{\alpha}}^2Q^2\vert \Phi(t)- W( \Phi(t-\delta))\vert^{2{\hat{\alpha}}-2}\times \left(1+\vert \Phi\vert^{\alpha_2}+\vert \Phi(t-\delta)\vert^{\alpha_2}\right)^2+\frac{\mu}{4\bar{c}}\bigg)dt. \end{aligned} |
For any 1\leq\bar{p}\leq \alpha , we get
\mathbb{E}\vert \Phi(t+s)\vert^{\bar{p}}\leq 1+\mathbb{E}\vert \Phi(t+s)\vert^{\alpha}, |
which further leads to
\begin{aligned} \sup\limits_{-\delta\leq s < 0}\mathbb{E}\vert \Phi(t+s)\vert^{\bar{p}}&\leq 1+\sup\limits_{-\delta\leq s < 0}\mathbb{E}\vert \Phi(t+s)\vert^{\alpha}\\ &\leq 1+\sup\limits_{-\delta\leq t < \infty}\mathbb{E}\vert \Phi(t)\vert^{\alpha}. \end{aligned} |
Therefore, according to Theorem 3.1, it follows that
\begin{align} \begin{aligned} \mathbb{E}\vert \Phi(t-\delta)\vert^{\bar{p}}&\leq \sup\limits_{-\delta\leq s < 0}\mathbb{E}\vert \Phi(t+s)\vert^{\bar{p}}\\ &\leq 1+\sup\limits_{-\delta\leq t < \infty}\mathbb{E}\vert \Phi(t)\vert^{\alpha} < \infty. \end{aligned} \end{align} | (3.30) |
By applying the inequality
\begin{aligned} &\vert \Phi(t)- W( \Phi(t-\delta))\vert^{\hat{\alpha}}\leq 2^{{\hat{\alpha}}-1}(\vert \Phi(t)\vert^{\hat{\alpha}}+\vert W( \Phi(t-\delta))\vert^{\hat{\alpha}})\\ &\quad\quad\quad\quad\quad\quad\quad\quad\; \; \leq 2^{{\hat{\alpha}}-1}(\vert \Phi(t)\vert^{\hat{\alpha}}+ \tilde{\omega }^{\hat{\alpha}}\vert \Phi(t-\delta)\vert^{\hat{\alpha}}),\\ &\vert \Phi(t)\vert^{{\hat{\alpha}}-1}\leq 1+\vert \Phi(t)\vert^{\alpha},\\ &\vert \Phi(t)\vert^{2({\hat{\alpha}}-1)}\leq 1+\vert \Phi(t)\vert^{\alpha},\\ &\vert \Phi(t)\vert^{{\hat{\alpha}}-1}\vert \Phi(t-\delta )\vert^{\alpha_1}\leq \vert \Phi(t)\vert^{{\hat{\alpha}}+\alpha_1-1}+\vert \Phi(t-\delta)\vert^{{\hat{\alpha}}+\alpha_1-1},\\ &\vert \Phi(t)\vert^{2({\hat{\alpha}}-1)}\vert \Phi(t-\delta )\vert^{2\alpha_2}\leq \vert \Phi(t)\vert^{2({\hat{\alpha}}+\alpha_2-1)}+\vert \Phi(t-\delta)\vert^{2({\hat{\alpha}}+\alpha_2-1)}, \end{aligned} |
and (3.29), we can get
\begin{aligned} \vert\mathbb{E}& \vert \Phi(t_2)- W( \Phi(t_2-\delta))\vert^{\hat{\alpha}}-\mathbb{E}\vert \Phi(t_1)- W( \Phi(t_1-\delta))\vert^{\hat{\alpha}}\vert\\ \leq&\mathbb{E}\int_{t_1}^{t_2}\bigg[ \hat{\alpha}Q2^{\hat{\alpha}-2}\left(\vert\Phi(t)\vert^{\hat{\alpha}-1}+\tilde{\omega}^{\hat{\alpha}-1}\vert\Phi(t-\delta)\vert^{\hat{\alpha}-1}\right)\\ &\times\left(1+\vert\Phi(t)\vert^{\alpha_{1}}+\vert\Phi(t-\delta)\vert^{\alpha_{1}}\right)\\ &+\bar{c}\hat{\alpha}^2Q^22^{2\hat{\alpha}-4}\left(\vert\Phi(t)\vert^{\hat{\alpha}-1}+\tilde{\omega}^{\hat{\alpha}-1}\vert\Phi(t-\delta)\vert^{\hat{\alpha}-1}\right)^2\\ &\times\left(1+\vert\Phi(t)\vert^{\alpha_{2}}+\vert\Phi(t-\delta)\vert^{\alpha_{2}}\right)^2+\frac{\mu}{4\bar{c}}\bigg]dt\\ \leq&\mathbb{E}\int_{t_1}^{t_2}\bigg[2^{\hat{\alpha}+1}\hat{\alpha}Q(1+\sup\limits_{-\delta\leq t < \infty}\mathbb{E}\vert\Phi(t)\vert^{\alpha})+\bar{c}\hat{\alpha}^2Q^22^{2\hat{\alpha}+2}\\ &\times(1+\sup\limits_{-\delta\leq t < \infty}\mathbb{E}\vert\Phi(t)\vert^{\alpha})+\frac{\mu}{4\bar{c}}\bigg]dt\\ \leq& K_4(t_2-t_1), \end{aligned} |
where
\begin{aligned} K_4& = \frac{\mu}{4\bar{c}}+2^{{\hat{\alpha}}+1}\left[{\hat{\alpha}}Q+2^{{\hat{\alpha}}+1}\bar{c}{\hat{\alpha}}^2Q^2\right](1+\sup\limits_{-\delta\leq t < \infty}\mathbb{E}\vert \Phi(t)\vert^\alpha)\\ & < \infty. \end{aligned} |
As a consequence, \mathbb{E}\vert \Phi(t)- W(\Phi(t-\delta))\vert^{\hat{\alpha}} is uniformly continuous. Based on (3.29), one has
\begin{aligned} \int_{0}^{\infty}\mathbb{E}\vert \Phi(t)- W( \Phi(t-\delta))\vert^{\hat{\alpha}}dt&\leq\int_{0}^{\infty}2^{{\hat{\alpha}}-1}\mathbb{E}\left(\vert \Phi(t)\vert^{\hat{\alpha}}+ \tilde{\omega }^{\hat{\alpha}}\vert \Phi(t-\delta)\vert^{\hat{\alpha}}\right)dt\\ &\leq 2^{{\hat{\alpha}}-1}(1+ \tilde{\omega }^{\hat{\alpha}})\int_{0}^{\infty}\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}dt+2^{{\hat{\alpha}}-1} \tilde{\omega }^{\hat{\alpha}}\delta\vert\vert \eta\vert\vert < \infty, \end{aligned} |
applying the Barbalat lemma, we have \lim_{t\to\infty}\mathbb{E}\vert \Phi(t)- W(\Phi(t-\delta))\vert^{\hat{\alpha}} = 0 . Next, applying the following inequality
(m+n)^{\hat{\alpha}}\leq (1+\varepsilon)^{{\hat{\alpha}}-1}(m^{\hat{\alpha}}+\varepsilon ^{1-{\hat{\alpha}}}n^{\hat{\alpha}}), \quad \forall m,n\geq 0,\ {\hat{\alpha}}\geq 1,\ \varepsilon > 0, |
we derive
\begin{aligned} \mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}&\leq \mathbb{E}\left[\vert \Phi(t)- W( \Phi(t-\delta))\vert+\vert W( \Phi(t-\delta))\vert\right]^{\hat{\alpha}}\\ &\leq \mathbb{E}\bigg[(1+\varepsilon)^{{\hat{\alpha}}-1}(\vert \Phi(t)- W( \Phi(t-\delta))\vert^{\hat{\alpha}}+\varepsilon^{1-{\hat{\alpha}}} \tilde{\omega }^{\hat{\alpha}}\vert \Phi(t-\delta)\vert^{\hat{\alpha}})\bigg]. \end{aligned} |
Taking \varepsilon = \frac{ \tilde{\omega }}{1- \tilde{\omega }} ,
\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}\leq(\frac{1}{1- \tilde{\omega }})^{{\hat{\alpha}}-1}\mathbb{E}\vert \Phi(t)- W( \Phi(t-\delta))\vert^{\hat{\alpha}}+ \tilde{\omega }\mathbb{E}\vert \Phi(t-\delta)\vert^{\hat{\alpha}}. |
Then, letting t\to\infty , we obtain
\lim\limits_{t\to\infty}\sup\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}\leq \tilde{\omega }\lim\limits_{t\to\infty}\sup\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}\quad a.s. |
By (3.29), one obtains \lim_{t\to\infty}\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}} = 0 .
Theorem 3.4. If the conditions of Theorem 3.2 are met and there exist two positive constants d > 0 and \hat{\alpha} > 0 satisfying
\begin{align} d\vert \Phi\vert^{\hat{\alpha}}\leq U_1( \Phi),\quad \end{align} | (3.31) |
then the solution of hybrid NSDDE (2.1) is almost surely asymptotically stability, i.e., \lim_{t\to\infty} \Phi(t) = 0 a.s.
Proof. Combined with (3.18), (3.25), and (3.27), we get
\int_{0}^{\infty}\mathbb{E}G( \Phi(t)- W( \Phi(t-\delta)))dt < \infty. |
According to Fubini's theorem, we get
\begin{aligned} \mathbb{E}\int_{0}^{\infty}G( \Phi(t)- W( \Phi(t-\delta)))dt < \infty, \end{aligned} |
which means
\begin{align} \int_{0}^{\infty}G( \Phi(t)- W( \Phi(t-\delta)))dt < \infty\; \; a.s. \end{align} | (3.32) |
Setting \bar{ \Phi}(t) = \Phi(t)- W(\Phi(t-\delta)) for t\geq 0 and \rho_h = \inf\left\{t\geq 0:\vert \bar{ \Phi}(t)\vert = h\right\} , by (3.32),
\begin{align} \begin{aligned} \lim\limits_{t\to\infty}\inf G(\bar{ \Phi}(t)) = 0\quad a.s. \end{aligned} \end{align} | (3.33) |
According to Corollary 3.1, we denote K_5: = \int_{0}^{\infty}\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}dt < \infty . Then, the proof follows a similar process to that of Theorem 3.3, and we obtain that
\begin{aligned} \mathbb{E}\vert \bar{ \Phi}(T\wedge\rho_h)\vert^{\hat{\alpha}}&\leq K_6+K_7\int_{0}^{\infty}\mathbb{E}\vert \Phi(t)\vert^{\hat{\alpha}}dt\\ & = K_6+K_5K_7: = K, \quad \quad\forall T > 0, \end{aligned} |
where K_6 = 2^{{\hat{\alpha}}-1} \tilde{\omega }^{\hat{\alpha}}\delta\vert\vert\eta\vert\vert, \ K_7 = 2^{{\hat{\alpha}}-1}(1+ \tilde{\omega }^{\hat{\alpha}}) . This implies
h^{\hat{\alpha}}\mathbb{P}(\rho_h\leq T)\leq K. |
Letting T\to\infty , it follows that
\begin{align} \begin{aligned} h^{\hat{\alpha}}\mathbb{P}(\rho_h < \infty)\leq K. \end{aligned} \end{align} | (3.34) |
The remainder of the proof will be segmented into three steps. First, we assert that
\begin{align} \begin{aligned} \lim\limits_{t\to\infty}G(\bar{ \Phi}(t)) = 0\quad\quad a.s. \end{aligned} \end{align} | (3.35) |
If Eq (3.35) is not fulfilled, then a sufficiently small constant \epsilon\in(0, \frac{1}{4}) can be found which satisfies
\begin{align} \begin{aligned} \mathbb{P}(\Delta_1)\geq 4\epsilon, \end{aligned} \end{align} | (3.36) |
where \Delta_1 = \left\{\lim_{t\to\infty}\sup G(\bar{ \Phi}(t)) > 2\epsilon\right\} . From (3.34), there exists a sufficiently large constant l with \mathbb{P}(\rho_l < \infty)\leq\epsilon , which means that
\begin{align} \begin{aligned} \mathbb{P}(\Delta_2)\geq1-\epsilon, \end{aligned} \end{align} | (3.37) |
where \Delta_2 = \left\{\vert \bar{ \Phi}(t)\vert < l\ \text{for}\ \forall t\geq-\delta\right\} . From (3.36) and (3.37), we can obtain
\begin{align} \begin{aligned} \mathbb{P}(\Delta_1\cap\Delta_2)\geq\mathbb{P}(\Delta_1)-\mathbb{P}(\Delta_2^c)\geq 3\epsilon. \end{aligned} \end{align} | (3.38) |
For t\geq-\delta , let \varsigma(t) = \bar{ \Phi}(t\wedge\rho_l) . It is clear that \varsigma(t) is bounded and
\begin{align} \begin{aligned} d\varsigma(t) = \hat{f}(t)dt+\hat{ \sigma}(t) \xi(t)dt, \end{aligned} \end{align} | (3.39) |
where
\begin{aligned} &\hat{f}(t) = f( \Phi(t), \Phi(t-\delta), \pi(t),t)I_{[0,\rho_l)}(t),\\ &\hat{ \sigma}(t) = \sigma( \Phi(t), \Phi(t-\delta), \pi(t),t)I_{[0,\rho_l)}(t). \end{aligned} |
For 0\leq t < \rho_l , from (2.5), we can get
\begin{aligned} \vert \Phi(t)\vert&\leq\vert \Phi(t)- W( \Phi(t-\delta))\vert+\vert W( \Phi(t-\delta))\vert\\ &\leq l+ \tilde{\omega }\vert \Phi(t-\delta)\vert, \end{aligned} |
which indicates
\sup\limits_{0\leq t < \rho_l}\vert \Phi(t)\vert\leq l+ \tilde{\omega }\vert\vert\eta\vert\vert+ \tilde{\omega }\sup\limits_{0\leq t < \rho_l}\vert \Phi(t)\vert. |
Therefore, there holds
\begin{align} \begin{aligned} \sup\limits_{-\delta\leq t < \rho_l}\vert \Phi(t)\vert\leq\left(\frac{1}{1- \tilde{\omega }}(l+ \tilde{\omega }\vert\vert\eta\vert\vert)\right)\vee\vert\vert\eta\vert\vert. \end{aligned} \end{align} | (3.40) |
From Assumption 2.2 and (3.40), it can be seen that \hat{f}(t) and \hat{ \sigma}(t) are bounded processes, and
\begin{align} \begin{aligned} \vert\hat{f}(t)\vert\vee\vert\hat{ \sigma}(t)\vert\leq K_8\quad\quad a.s. \end{aligned} \end{align} | (3.41) |
where all t\geq 0 and some K_8 > 0 . From the definition of \rho_l , it is easy to get \vert\varsigma(t)\vert\leq l for any t\geq-\delta .
Set the stopping time
\begin{aligned} \psi_1& = \inf\left\{t\geq 0:G(\varsigma(t))\geq 2\epsilon\right\},\\ \psi_{2q}& = \inf\left\{t\geq \psi_{2q-1}:G(\varsigma(t))\leq \epsilon\right\},\quad q = 1,2,\dots,\\ \psi_{2q+1}& = \inf\left\{t\geq \psi_{2q}:G(\varsigma(t))\geq 2\epsilon\right\},\quad q = 1,2,\dots. \end{aligned} |
Based on (3.33), as well as the definitions of \Delta_1 and \Delta_2 , it follows that
\begin{align} \begin{aligned} \Delta_1\cap\Delta_2\subset\left\{\rho_{l} = \infty\right\}\cap\left(\cap_{q = 1}^{\infty}\left\{\psi_q < \infty\right\}\right). \end{aligned} \end{align} | (3.42) |
For all \omega\in\Delta_1\cap\Delta_2 and q\geq 1 , there are
\begin{align} \begin{aligned} &G(\varsigma(\psi_{2q-1}))-G(\varsigma(\psi_{2q})) = \epsilon\quad \text{and}\quad\\ &G(\varsigma(t))\geq\epsilon,\quad t\in[\psi_{2q-1},\psi_{2q}]. \end{aligned} \end{align} | (3.43) |
We know that G(\cdot) is uniformly continuous in \bar{S}_l = \left\{ \Phi\in \mathcal{R}^n:\vert \Phi\vert\leq l\right\} . It is possible to find \tau = \tau(\epsilon) > 0 small enough to make
\begin{align} \begin{aligned} \vert G(\varsigma_1)-G(\varsigma_2)\vert < \epsilon,\; \varsigma_1,\; \varsigma_1\in\bar{S}_l,\ \text{with} \ \ \vert \varsigma_1-\varsigma_2\vert < \tau.\quad\quad\quad \end{aligned} \end{align} | (3.44) |
We highlight that, for \omega\in\Delta_1\cap\Delta_2 , if \vert\varsigma(\psi_{2q-1}+v)-\varsigma(\psi_{2q-1})\vert < \tau for all v\in[0, \Upsilon ] and some \Upsilon > 0 , then \psi_{2q}-\psi_{2q-1}\geq\Upsilon . Accordingly, there exists a small enough constant \Upsilon > 0 and a large enough integer q_0 > 0 such that
\begin{align} \begin{aligned} 2K_8^2\Upsilon ^2(1+\mu)\leq\epsilon\tau^2\quad\text{and}\quad\mathbb{E}\int_{0}^{\infty}G(\bar{ \Phi}(t))dt < \epsilon^2\Upsilon q_0. \end{aligned} \end{align} | (3.45) |
By (3.38) and (3.42), there exists a constant T large enough such that
\begin{align} \begin{aligned} \mathbb{P}(\psi_{2q_0}\leq T)\geq 2\epsilon. \end{aligned} \end{align} | (3.46) |
If \psi_{2q_0}\leq T , then \vert\varsigma(\psi_{2q_0})\vert < l , and thus \psi_{2q_0} < \rho_l . So, for any 0\leq t\leq\psi_{2q_0} , as well as \omega\in\left\{\psi_{2q_0}\leq T\right\} , there holds
\begin{align} \varsigma(t,\omega) = \bar{ \Phi}(t,\omega). \end{align} | (3.47) |
Together with Assumption 3.1 and (3.41), for 1\leq q\leq q_0 , we obtain
\begin{align} \begin{aligned} \mathbb{E}&\left(\sup\limits_{0\leq t\leq\Upsilon }\vert\varsigma(\psi_{2q-1}\wedge T+t)-\varsigma(\psi_{2q-1}\wedge T)\vert^2\right)\\ &\leq 2\Upsilon \mathbb{E}\int_{\psi_{2q-1}\wedge T+\Upsilon }^{\psi_{2q-1}\wedge T}\vert\hat{f}(s)\vert^2ds+2\mu\Upsilon \mathbb{E}\int_{\psi_{2q-1}\wedge T+\Upsilon }^{\psi_{2q-1}\wedge T}\vert\hat{ \sigma}(s)\vert^2ds\\ &\leq2K_8^2\Upsilon^2(1+\mu). \end{aligned} \end{align} | (3.48) |
Based on the Chebyshev inequality and (3.45), there holds
\begin{align} \mathbb{P}\left(\sup\limits_{0\leq t\leq\Upsilon }\vert\varsigma(\psi_{2q-1}\wedge T+t)-\varsigma(\psi_{2q-1}\wedge T)\vert\geq\tau\right)\leq\epsilon. \end{align} | (3.49) |
If \psi_{2q_0}\leq T , then \psi_{2q-1}\leq T , and combining (3.46) and (3.49) yields
\begin{aligned} \mathbb{P}&\left(\left\{\psi_{2q_0}\leq T\right\}\cap\left\{\sup\limits_{0\leq t\leq\Upsilon }\vert\varsigma(\psi_{2q-1}+t)-\varsigma(\psi_{2q-1})\vert < \tau\right\}\right)\\ = &\mathbb{P}(\psi_{2q_0}\leq T)-\mathbb{P}\left(\left\{\psi_{2q_0}\leq T\right\}\cap\left\{\sup\limits_{0\leq t\leq\Upsilon }\vert\varsigma(\psi_{2q-1}+t)-\varsigma(\psi_{2q-1})\vert\geq\tau\right\}\right)\\ \geq&\mathbb{P}(\psi_{2q_0}\leq T)-\mathbb{P}\left(\sup\limits_{0\leq t\leq\Upsilon }\vert\varsigma(\psi_{2q-1}+t)-\varsigma(\psi_{2q-1})\vert\geq\tau\right)\\ \geq&\epsilon. \end{aligned} |
Based on (3.44), this implies that
\begin{align} \mathbb{P}\left(\left\{\psi_{2q_0}\leq T\right\}\cap\left\{\psi_{2q}-\psi_{2q-1}\geq\Upsilon \right\}\right)\geq\epsilon. \end{align} | (3.50) |
By (3.32), (3.47), and (3.50), we conclude
\begin{aligned} \mathbb{E}\int_{0}^{\infty}G(\bar{ \Phi}(t))dt&\geq\sum\limits_{q = 1}^{q_0}\mathbb{E}\left(I_{\left\{\psi_{2q_0}\leq T\right\}}\int_{\psi_{2q-1}}^{\psi_{2q}}G(\bar{ \Phi}(t))dt\right)\\ &\geq\epsilon\sum\limits_{q = 1}^{q_0}\mathbb{E}\left(I_{\left\{\psi_{2q_0}\leq T\right\}}(\psi_{2q}-\psi_{2q-1})\right)\\ &\geq\epsilon\Upsilon \sum\limits_{q = 1}^{q_0}\mathbb{P}\left(\left\{\psi_{2q_0}\leq T\right\}\cap\left\{\psi_{2q}-\psi_{2q-1}\geq\Upsilon \right\}\right)\\ &\geq\epsilon^2\Upsilon q_0, \end{aligned} |
which conflicts with (3.45). Thus, (3.35) must hold.
The second step involves proving that
\lim\limits_{t\to\infty}\bar{ \Phi}(t) = 0\quad a.s. |
If this is false, then \epsilon_0 = \mathbb{P}(\Delta_3) > 0 , where \Delta_3 = \left\{\limsup_{t\to\infty}\vert \bar{ \Phi}(t)\vert > 0\right\} . By (3.34), there exists a large enough integer m_0 > 0 such that \mathbb{P}(\rho_{m_0} < \infty)\leq\frac{1}{2}\epsilon_0 . Let \Delta_4 = \left\{\rho_{m_0} = \infty\right\} . Then,
\mathbb{P}(\Delta_3\cap\Delta_4)\geq\mathbb{P}(\Delta_3)-\mathbb{P}(\Delta_4^c)\geq\frac{1}{2}\epsilon_0. |
Note that, for any \omega\in\Delta_3\cap\Delta_4 and t\geq 0 , \bar{ \Phi}(t, \omega) is bounded. It is possible to find a sequence \left\{t_i\right\}_{i\geq 1} satisfying t_i\to\infty as well as \bar{ \Phi}(t_i, \omega)\to \bar{ \Phi}(\omega)\ne0 as i\to\infty . It is worth noting that, since G is continuous, we can obtain
\lim\limits_{j\to\infty}G(\bar{ \Phi}(t_i,\omega)) = G(\bar{ \Phi}(\omega)) > 0. |
Therefore, for all \omega\in\Delta_3\cap\Delta_4 ,
\limsup\limits_{t\to\infty}G(\bar{ \Phi}(t,\omega)) > 0. |
But, this contradicts (3.35). Thus, we can obtain \lim_{t\to\infty}\bar{ \Phi}(t) = 0 a.s. Further, we can get
\begin{align} \sup\limits_{0\leq t < \infty}\vert \bar{ \Phi}(t)\vert < \infty\quad a.s. \end{align} | (3.51) |
The third step involves claiming assertion (3.31). It follows from (2.5) that
\begin{align} \begin{aligned} \vert \Phi(t)\vert &\leq \vert \Phi(t)- W( \Phi(t-\delta))\vert +\vert W( \Phi(t-\delta))\vert\\ &\leq \vert \bar{ \Phi}(t)\vert + \tilde{\omega }\vert \Phi(t-\delta)\vert \quad a.s. \end{aligned} \end{align} | (3.52) |
Then, for any T > 0 ,
\sup\limits_{0\leq t < T}\vert \Phi(t)\vert\leq \sup\limits_{0\leq t < T}\vert \bar{ \Phi}(t)\vert+ \tilde{\omega }\vert\vert\eta\vert\vert+ \tilde{\omega }\sup\limits_{0\leq t < T}\vert \Phi(t)\vert\quad a.s. |
Consequently, we have
\sup\limits_{0\leq t < T}\vert \Phi(t)\vert\leq\frac{1}{1- \tilde{\omega }}\left(\sup\limits_{0\leq t < T}\vert \bar{ \Phi}(t)\vert+ \tilde{\omega }\vert\vert\eta\vert\vert\right)\quad a.s. |
Making use of (3.51) and allowing T\to\infty , we get
\begin{align} \sup\limits_{0\leq t < \infty}\vert \Phi(t)\vert < \infty\quad a.s. \end{align} | (3.53) |
Letting t\to\infty in (3.52) and combining \lim_{t\to\infty}\bar{ \Phi}(t) = 0 a.s., we have
\limsup\limits_{t\to\infty}\vert \Phi(t)\vert\leq \tilde{\omega }\limsup\limits_{t\to\infty}\vert \Phi(t)\vert\quad a.s. |
Since \tilde{\omega }\in(0, \frac{\sqrt{2}}{2}) , and by (3.53), we obtain
\lim\limits_{t\to\infty}\vert \Phi(t)\vert = 0\quad a.s. |
Remark 3.4. When the noise considered in hybrid NSDDE (2.1) is white noise, we obtain that Theorems 3.2–3.4 that are consistent with those in [17].
Remark 3.5. In contrast to [10], in this paper, we develop new mathematical techniques to address the challenges posed by the neutral term, since the presence of the neutral term fundamentally alters the issue.
Remark 3.6. The nonlinear functions considered in [24] satisfy the linear growth condition. When \alpha_1 = \alpha_2 = 1 in Assumption 2.2, the PGC simplifies to the LGC, and thus the nonlinear functions under consideration throughout the paper are more universal.
We will validate the correctness of the theoretical results through examples in this section.
Let us examine the highly nonlinear hybrid NSDDE with colored noise (1.1). Based on the coefficients of (1.1), Assumptions 2.1–2.3, and 3.1 hold when Q = 6 , \alpha_1 = 3 , \alpha_2 = 1 , \tilde{\omega } = 0.1 and \mu = 0.15 . Let \bar{U}(\Phi, t, i) = \vert \Phi\vert^6 . Then, we get
\begin{aligned} d\bar{U}(\bar \Phi,\nu,i,t)&\leq\left\{\begin{aligned} &-20.0248 \Phi^8+11.5968 \Phi^6+2.46 \nu^8+3.5811 \nu^6+0.025\vert \xi(t)\vert^{24}, &i = 1,\\ &-20.0247 \Phi^8+8.5525 \Phi^6+2.535 \nu^8+2.4771 \nu^6+0.05\vert \xi(t)\vert^{24}, &i = 2, \end{aligned} \right. \end{aligned} |
which shows
\begin{aligned} d\bar{U}(\bar{ \Phi}, \nu,i,t)&\leq15.2968-3.7( \Phi^8+ \Phi^6)+3.5811( \nu^8+ \nu^6)+0.05\vert \xi(t)\vert^{24}\\ &\leq b_{1}-b_{2}H( \Phi)+b_{3}H( \nu)+\gamma(\vert \xi(t)\vert^{2}), \end{aligned} |
where b_1 = 15.2968, \ b_2 = 3.7, \ b_3 = 3.5811, \bar{ \Phi} = \Phi-0.1 \nu, \ H(\Phi) = \Phi^8+ \Phi^6 , and \gamma(\vert \xi(t)\vert^2) = 0.05\vert \xi(t)\vert^{24}. Hence, it can be concluded that Assumption 2.4 is also fulfilled.
Define the function as follows:
U( \Phi,i,t) = \left\{\begin{aligned} &\frac{1}{2} \Phi^2+\frac{1}{4} \Phi^4,&i = 1,\\ &\frac{3}{4} \Phi^2+\frac{1}{4} \Phi^4,&i = 2. \end{aligned} \right. |
By calculating, we get
\vert U_ \Phi(\bar{ \Phi},i,t)\vert^2\leq\left\{ \begin{aligned} &\bar{ \Phi}^2+2\bar{ \Phi}^4+\bar{ \Phi}^6,&i = 1,\\ &\frac{9}{4}\bar{ \Phi}^2+3\bar{ \Phi}^4+\bar{ \Phi}^6,&i = 2. \end{aligned} \right. |
From (3.12), we get
\begin{aligned} \mathcal{L}U(\bar{ \Phi}, \nu,1,t)\leq&-3.6917 \Phi^6-6.0466 \Phi^4-1.205 \Phi^2\\ &+0.3851 \nu^6+0.3453 \nu^4+0.8769 \nu^2\\ &-0.5\bar{ \Phi}^6-0.5\bar{ \Phi}^4-0.2\bar{ \Phi}^2\\ &+0.05 \Phi^6\vert \xi(t)\vert^2+0.015 \Phi^4\vert \xi(t)\vert^2+0.0515 \Phi^2\vert \xi(t)\vert^2 \end{aligned} |
and
\begin{aligned} \mathcal{L}U(\bar{ \Phi}, \nu,2,t)\leq&-4.0138 \Phi^6-8.4342 \Phi^4-1.095 \Phi^2\\ &+0.3543 \nu^6+0.3975 \nu^4+1.048 \nu^2\\ &-0.3\bar{ \Phi}^6-0.5\bar{ \Phi}^4-0.3\bar{ \Phi}^2\\ &+0.1 \Phi^6\vert \xi(t)\vert^2+0.03 \Phi^4\vert \xi(t)\vert^2+0.153 \Phi^2\vert \xi(t)\vert^2. \end{aligned} |
Moreover,
\begin{aligned} &\vert f( \Phi, \nu,i,t)\vert^2\leq\left\{ \begin{aligned} &72 \Phi^6+4.5 \nu^2,&i = 1,\\ &72 \Phi^6+2 \nu^2,&i = 2, \end{aligned} \right.\\ &\vert \sigma( \Phi, \nu,i,t)\vert^2\leq\left\{ \begin{aligned} &0.01 \nu^2,&i = 1,\\ &0.04 \nu^2,&i = 2. \end{aligned} \right. \end{aligned} |
Choosing \vartheta _{1} = 0.1 and \vartheta _{2} = 0.01, we have
\begin{aligned} \mathcal{L}&U(\bar{ \Phi},\nu,1,t)+\vartheta _{1}\vert U_{ \Phi}(\bar{ \Phi},1,t)\vert^{2}+\vartheta _{2}\vert f( \Phi, \nu,1,t)\vert^{2}+\mu\vartheta _{2}\vert \sigma( \Phi, \nu,1,t)\vert^{2}\\ \leq&-2.9717 \Phi^6-6.0466 \Phi^4-1.205 \Phi^2+0.3851 \nu^6+0.3453 \nu^4+0.9219 \nu^2\\ &-0.4\bar{ \Phi}^6-0.3\bar{ \Phi}^4-0.1\bar{ \Phi}^2+0.05 \Phi^6\vert \xi(t)\vert^2+0.015 \Phi^4\vert \xi(t)\vert^2+0.0515 \Phi^2\vert \xi(t)\vert^2 \end{aligned} |
and
\begin{aligned} \mathcal{L}&U(\bar{ \Phi},\nu,2,t)+\vartheta _{1}\vert U_{ \Phi}(\bar{ \Phi},2,t)\vert^{2}+\vartheta _{2}\vert f( \Phi, \nu,2,t)\vert^{2}+\mu\vartheta _{2}\vert \sigma( \Phi, \nu,2,t)\vert^{2}\\ \leq&-3.2938 \Phi^6-8.4342 \Phi^4-1.095 \Phi^2+0.3543 \nu^6+0.3975 \nu^4+1.0681 \nu^2\\ &-0.2\bar{ \Phi}^6-0.2\bar{ \Phi}^4-0.075\bar{ \Phi}^2+0.1 \Phi^6\vert \xi(t)\vert^2+0.03 \Phi^4\vert \xi(t)\vert^2+0.153 \Phi^2\vert \xi(t)\vert^2. \end{aligned} |
Thus,
\begin{aligned} &\mathcal{L}U(\bar{ \Phi},\nu,i,t)+\vartheta _{1}\vert U_{ \Phi}(\bar{ \Phi},i,t)\vert^{2}+\vartheta _{2}\vert f( \Phi, \nu,i,t)\vert^{2}+\mu\vartheta _{2}\vert \sigma( \Phi, \nu,i,t)\vert^{2}\\ &\leq-1.905( \Phi^6+ \Phi^4+ \Phi^2)+1.0681( \nu^6+ \nu^4+ \nu^2)\\ &\quad+0.153( \Phi^6+ \Phi^4+ \Phi^2)\vert \xi(t)\vert^2-0.2\bar{ \Phi}^6-0.2\bar{ \Phi}^4-0.075\bar{ \Phi}^2. \end{aligned} |
Let \beta_{1} = 1.905 , \beta_{2} = 1.0681 , \beta_{3} = 0.153 , U_1(\Phi) = \Phi^6+ \Phi^4+ \Phi^2 , and G(\Phi) = 0.2 \Phi^6 +0.2 \Phi^4+0.075 \Phi^2. It is easy to demonstrate that Assumptions 3.2 and 3.3, along with condition (3.16) when L = 1.5 , have been satisfied. Thus, by condition (3.17), we have \delta\leq 0.0209 . Moreover, according to Theorem 3.2, the unique global solution of (1.1) satisfies both (3.18) and (3.19). For \hat{\alpha}\in[2, 6], \ d = 1 , by Corollary 3.1, we can get (1.1) is H_{\infty} -stable in L^{\hat{\alpha}} . Since \alpha_1 = 3, \alpha_2 = 1 , and \alpha = 6 , for \hat{\alpha} = 3 , by Theorems 3.3 and 3.4, it follows that the global solution of (1.1) is asymptotically stable in L^{\hat{\alpha}} and almost surely asymptotically stable. We show a computer simulation of (1.1) with \delta = 0.02 in Figure 3. It is obvious from Figure 3 that the global solution of (1.1) is stable.
Remark 4.1. Literature [24] derived NSS criteria for neutral stochastic delayed nonlinear systems, however the impact of Markov switching was not considered. On the other hand, literature [22] investigated the DDS of a class of multi-delay hybrid neutral SDEs, but the influence of colored noise was not addressed. Building upon these studies, this paper incorporates both Markov switching and colored noise to develop a more comprehensive stability analysis framework.
Remark 4.2. Hybrid NSDDEs with colored noise form a class of mathematical tools that can efficiently model complex dynamical systems, and are especially suitable for describing systems with stochastic, nonlinear, time delay, and Markov switching properties. In addition to power systems, hybrid NSDDEs with colored noise have applications in other areas. For example, in robotic arm motion control, hybrid NSDDEs can be used to optimize trajectory tracking performance and improve control accuracy. In finance, they can be used to model the dynamic behavior of stock prices and predict their future trends. By considering these complex factors, hybrid NSDDEs can more accurately portray the dynamic characteristics of real systems and provide strong theoretical support for system analysis and control.
The existence of global solution of highly nonlinear hybrid NSDDEs has been proved under PGC, and the NSS- \alpha -M of the global solution has been obtained by inequality techniques. Furthermore, the Lyapunov function method was utilized to construct several innovative DDS criteria for highly nonlinear hybrid NSDDEs, including H_{\infty} -stability in L^{\hat{\alpha}} , asymptotic stability in L^{\hat{\alpha}} , and almost surely asymptotic stability. In future work, we will investigate highly nonlinear hybrid NSDDEs with multiple time delays or L \acute{\text{e}} vy noise [26], and explore the application of highly nonlinear hybrid NSDDEs to biological models [27,28].
Siru Li: Writing-original draft; Tian Xu: Writing-review and editing; Ailong Wu: Supervision, writing-review and editing. All authors have read and approved the final version of the manuscript for publication.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this paper.
This work is supported by the National Natural Science Foundation of China under Grant 62476082.
The authors declare that there are no conflicts of interest.
[1] | E. Allen, Modeling with Itô Stochastic Differential Equations, Dordrecht: Springer, 2007. https://doi.org/10.1007/978-1-4020-5953-7 |
[2] |
M. X. Shen, C. Fei, W. Y. Fei, X. R. Mao, Stabilisation by delay feedback control for highly nonlinear neutral stochastic differential equations, Syst. Control Lett., 137 (2020), 104645. https://doi.org/10.1016/j.sysconle.2020.104645 doi: 10.1016/j.sysconle.2020.104645
![]() |
[3] |
Q. Luo, X. R. Mao, Y. Shen, New criteria on exponential stability of neutral stochastic differential delay equations, Syst. Control Lett., 55 (2006), 826–834. https://doi.org/10.1016/j.sysconle.2006.04.005 doi: 10.1016/j.sysconle.2006.04.005
![]() |
[4] |
Y. Xu, Z. M. He, Exponential stability of neutral stochastic delay differential equations with Markovian switching, Appl. Math. Lett., 52 (2016), 64–73. https://doi.org/10.1016/j.aml.2015.08.019 doi: 10.1016/j.aml.2015.08.019
![]() |
[5] |
B. L. Lu, Q. X. Zhu, P. He, Exponential stability of highly nonlinear hybrid differently structured neutral stochastic differential equations with unbounded delays, Fractal. Fract., 6 (2022), 385. https://doi.org/10.3390/fractalfract6070385 doi: 10.3390/fractalfract6070385
![]() |
[6] | X. R. Mao, Stochastic Differential Equations and Applications, Amsterdam: Elsevier, 2007. https://doi.org/10.1016/C2013-0-07332-X |
[7] | R. Khasminskii, Stochastic Stability of Differential Equations, Berlin: Springer, 2012,385. https://doi.org/10.1007/978-3-642-23280-0 |
[8] |
X. Y. Li, X. R. Mao, Stabilisation of highly nonlinear hybrid stochastic differential delay equations by delay feedback control, Automatica, 112 (2020), 108657. https://doi.org/10.1016/j.automatica.2019.108657 doi: 10.1016/j.automatica.2019.108657
![]() |
[9] |
S. Mekki, T. Blouhi, J. J. Nieto, A. Ouahab, Some existence results for systems of impulsive stochastic differential equations, Ann. Math. Siles., 35 (2021), 260–281. https://doi.org/10.2478/amsil-2020-0028 doi: 10.2478/amsil-2020-0028
![]() |
[10] |
N. Yang, J. Hu, D. Y. Chen, Noise-to-state and delay-dependent stability of highly nonlinear hybrid stochastic differential equation with multiple delays, Int. J. Control, 97 (2024), 577–588. https://doi.org/10.1080/00207179.2022.2160824 doi: 10.1080/00207179.2022.2160824
![]() |
[11] |
L. K. Feng, W. H. Zhang, Z. J. Wu, Noise-to-state stability of random impulsive delay systems with multiple random impulses, Appl. Math. Comput., 436 (2023), 127517. https://doi.org/10.1016/j.amc.2022.127517 doi: 10.1016/j.amc.2022.127517
![]() |
[12] |
R. L. Song, B. Wang, Q. X. Zhu, Stabilization by variable-delay feedback control for highly nonlinear hybrid stochastic differential delay equations, Syst. Control Lett., 157 (2021), 105041. https://doi.org/10.1016/j.sysconle.2021.105041 doi: 10.1016/j.sysconle.2021.105041
![]() |
[13] |
C. Fei, W. Y. Fei, L. T. Yan, Existence and stability of solutions to highly nonlinear stochastic differential delay equations driven by G-Brownian motion, Appl. Math. - J. Chinese Univ., 34 (2019), 184–204. https://doi.org/10.1007/s11766-019-3619-x doi: 10.1007/s11766-019-3619-x
![]() |
[14] |
Y. Zhao, Q. X. Zhu, Stabilization of stochastic highly nonlinear delay systems with neutral term, IEEE Trans. Autom. Control, 68 (2022), 2544–2551. https://doi.org/10.1109/TAC.2022.3186827 doi: 10.1109/TAC.2022.3186827
![]() |
[15] |
M. X. Shen, W. Y. Fei, X. R. Mao, S. N. Deng, Exponential stability of highly nonlinear neutral pantograph stochastic differential equations, Asian J. Control, 22 (2020), 436–448. https://doi.org/10.1002/asjc.1903 doi: 10.1002/asjc.1903
![]() |
[16] |
C. Fei, W. Y. Fei, X. R. Mao, M. X. Shen, L. T. Yan, Stability analysis of highly nonlinear hybrid multiple-delay stochastic differential equations, J. Appl. Anal. Comput., 9 (2019), 1053–1070. https://doi.org/ 10.11948/2156-907X.20180257 doi: 10.11948/2156-907X.20180257
![]() |
[17] |
M. X. Shen, W. Y. Fei, X. R. Mao, Y. Liang, Stability of highly nonlinear neutral stochastic differential delay equations, Syst. Control Lett., 115 (2018), 1–8. https://doi.org/10.1016/j.sysconle.2018.02.013 doi: 10.1016/j.sysconle.2018.02.013
![]() |
[18] |
W. Y. Fei, L. J. Hu, X. R. Mao, M. X. Shen, Generalized criteria on delay‐dependent stability of highly nonlinear hybrid stochastic systems, Int. J. Robust Nonlinear Control, 29 (2019), 1201–1215. https://doi.org/10.1002/rnc.4402 doi: 10.1002/rnc.4402
![]() |
[19] |
C. Fei, W. Y. Fei, X. R. Mao, L. T. Yan, Delay-dependent asymptotic stability of highly nonlinear stochastic differential delay equations driven by G-Brownian motion, J. Franklin Inst., 359 (2022), 4366–4392. https://doi.org/10.1016/j.jfranklin.2022.03.027 doi: 10.1016/j.jfranklin.2022.03.027
![]() |
[20] |
A. Zouine, H. Bouzahir, C. Imzegouan, Delay-dependent stability of highly nonlinear hybrid stochastic systems with Lévy noise, Nonlinear Stud., 27 (2020), 879–896. https://doi.org/10.1016/j.jfranklin.2022.03.027 doi: 10.1016/j.jfranklin.2022.03.027
![]() |
[21] |
W. Y. Fei, L. J. Hu, X. R. Mao, M. X. Shen, Delay dependent stability of highly nonlinear hybrid stochastic systems, Automatica, 82 (2017), 165–170. https://doi.org/10.1016/j.automatica.2017.04.050 doi: 10.1016/j.automatica.2017.04.050
![]() |
[22] |
R. L. Song, B. Wang, Q. X. Zhu, Delay-dependent stability of nonlinear hybrid neutral stochastic differential equations with multiple delays, Int. J. Robust Nonlinear Control, 31 (2021), 250–267. https://doi.org/10.1002/rnc.5275 doi: 10.1002/rnc.5275
![]() |
[23] |
M. X. Shen, C. Fei, W. Y. Fei, X. R. Mao, C. H. Mei, Delay‐dependent stability of highly nonlinear neutral stochastic functional differential equations, Int. J. Robust Nonlinear Control, 32 (2022), 9957–9976. https://doi.org/10.1002/rnc.6384 doi: 10.1002/rnc.6384
![]() |
[24] |
W. Lu, Noise-to-state exponential stability of neutral random nonlinear systems, Adv. Differ. Equ., 2018 (2018), 313. https://doi.org/10.1186/s13662-018-1739-z doi: 10.1186/s13662-018-1739-z
![]() |
[25] |
Z. J. Wu, Stability criteria of random nonlinear systems and their applications, IEEE Trans. Autom. Control, 60 (2014), 1038–1049. https://doi.org/10.1109/TAC.2014.2365684 doi: 10.1109/TAC.2014.2365684
![]() |
[26] |
H. L. Dong, J. Tang, X. R. Mao, Stabilization of highly nonlinear hybrid stochastic differential delay equations with Lévy noise by delay feedback control, SIAM J. Control Optim., 60 (2022), 3302–3325. https://doi.org/10.1137/22M1480392 doi: 10.1137/22M1480392
![]() |
[27] |
F. A. Rihan, H. J. Alsakaji, Stochastic delay differential equations of three-species prey-predator system with cooperation among prey species, Discrete Cont. Dyn. Syst. Ser. S., 15 (2022), 245–263. https://doi.org/10.3934/dcdss.2020468 doi: 10.3934/dcdss.2020468
![]() |
[28] |
F. A. Rihan, H. J. Alsakaji, S. Kundu, O. Mohamed, Dynamics of a time-delay differential model for tumour-immune interactions with random noise, Alex. Eng. J., 61 (2022), 11913–11923. https://doi.org/10.1016/j.aej.2022.05.027 doi: 10.1016/j.aej.2022.05.027
![]() |