
This research considers discontinuous dynamical systems, which have related vector fields that shift over a discontinuity surface. These systems appear in a variety of applications, including ecology, medicine, neuroscience, and nonsmooth mechanics. The purpose of this paper is to develop a perturbation technique that measures the effect of a nonsmooth perturbation on the period doubling bifurcation of an unperturbed system. The unperturbed system is assumed to be close to a period doubling orbit, such that when the bifurcation parameter varies, the response changes from a period one to a period two limit cycle. The generalized determination of the Poincaré map associated with perturbed systems subjected to nonsmooth transitions is derived. The main techniques used in the proof of the results are normal forms and Melnikov functions, which are defined in two zones. Various examples are presented to show that non-smoothness is responsible for period doubling. To illustrate the interesting period doubling phenomenon that emerges from an existing flat periodic orbit via the non-smooth perturbation, a simple and novel discontinuous system is provided. An additional example is provided to show the emergence of a perturbed period doubling orbit near an unperturbed one.
Citation: Hany A. Hosham, Alaa A. Alzulaibani, Tarek Sellami, Khaled Sioud, Thoraya N. Alharthi. A class of discontinuous systems exhibit perturbed period doubling bifurcation[J]. AIMS Mathematics, 2024, 9(9): 25098-25113. doi: 10.3934/math.20241223
[1] | Hany A. Hosham, Thoraya N. Alharthi . Bifurcation and chaos in simple discontinuous systems separated by a hypersurface. AIMS Mathematics, 2024, 9(7): 17025-17038. doi: 10.3934/math.2024826 |
[2] | Ninghe Yang . Exact wave patterns and chaotic dynamical behaviors of the extended (3+1)-dimensional NLSE. AIMS Mathematics, 2024, 9(11): 31274-31294. doi: 10.3934/math.20241508 |
[3] | Thoraya N. Alharthi . Dynamic analysis and multistability of a discontinuous Jerk-like system. AIMS Mathematics, 2025, 10(5): 12554-12575. doi: 10.3934/math.2025566 |
[4] | Limei Liu, Xitong Zhong . Analysis and anti-control of period-doubling bifurcation for the one-dimensional discrete system with three parameters and a square term. AIMS Mathematics, 2025, 10(2): 3227-3250. doi: 10.3934/math.2025150 |
[5] | M. Chandru, T. Prabha, V. Shanthi, H. Ramos . An almost second order uniformly convergent method for a two-parameter singularly perturbed problem with a discontinuous convection coefficient and source term. AIMS Mathematics, 2024, 9(9): 24998-25027. doi: 10.3934/math.20241219 |
[6] | Abdul Qadeer Khan, Ayesha Yaqoob, Ateq Alsaadi . Discrete Hepatitis C virus model with local dynamics, chaos and bifurcations. AIMS Mathematics, 2024, 9(10): 28643-28670. doi: 10.3934/math.20241390 |
[7] | Minzhi Wei, Liping He . Existence of periodic wave for a perturbed MEW equation. AIMS Mathematics, 2023, 8(5): 11557-11571. doi: 10.3934/math.2023585 |
[8] | Jia Li, Xia Li, Chunpeng Zhu . Reducibility for a class of almost periodic Hamiltonian systems which are degenerate. AIMS Mathematics, 2023, 8(1): 2296-2307. doi: 10.3934/math.2023119 |
[9] | Yanshou Dong, Junfang Zhao, Xu Miao, Ming Kang . Piecewise pseudo almost periodic solutions of interval general BAM neural networks with mixed time-varying delays and impulsive perturbations. AIMS Mathematics, 2023, 8(9): 21828-21855. doi: 10.3934/math.20231113 |
[10] | Yunsoo Jang . Global gradient estimates in directional homogenization. AIMS Mathematics, 2023, 8(11): 27643-27658. doi: 10.3934/math.20231414 |
This research considers discontinuous dynamical systems, which have related vector fields that shift over a discontinuity surface. These systems appear in a variety of applications, including ecology, medicine, neuroscience, and nonsmooth mechanics. The purpose of this paper is to develop a perturbation technique that measures the effect of a nonsmooth perturbation on the period doubling bifurcation of an unperturbed system. The unperturbed system is assumed to be close to a period doubling orbit, such that when the bifurcation parameter varies, the response changes from a period one to a period two limit cycle. The generalized determination of the Poincaré map associated with perturbed systems subjected to nonsmooth transitions is derived. The main techniques used in the proof of the results are normal forms and Melnikov functions, which are defined in two zones. Various examples are presented to show that non-smoothness is responsible for period doubling. To illustrate the interesting period doubling phenomenon that emerges from an existing flat periodic orbit via the non-smooth perturbation, a simple and novel discontinuous system is provided. An additional example is provided to show the emergence of a perturbed period doubling orbit near an unperturbed one.
In the field of physics, there has been significant attention given to disturbances affecting a nonlinear dynamic system that approaches a period doubling bifurcation (PDB). This objective has been studied in such a way that the systems whose equations are either discrete or continuous. For instance, in Josephson-junction amplifiers, the basic amplifying mechanism depends on the PDB [1]. Recently, it has been shown that the PDB has a significant effect on subharmonics and ultra-harmonic emissions of bubble oscillations [2]. Further, the PDB is a type of bifurcation commonly observed in many systems corresponding to applications from different fields of science, such as biological, chemical, electrical, hydrodynamical, and optical systems, for instances see [3,4,5].
The differential equations with discontinuous vector fields may show complicated nonlinear phenomena, such as, hidden and nonclassical various types of bifurcations and chaos, which may occur in practical mechanical models [6,7]. Therefore, the exploration of novel mathematical approaches for analyzing discontinuous systems is intriguing, yet challenging. This is because traditional methods like Lyapunov-Schmidt reduction and center manifold theory are no longer directly suitable for addressing sudden alterations in the governing vector fields. There has been a lot of progress in improving the traditional methods of bifurcations to study and classify singularities, attractors, and hidden chaotic attractors in discontinuous systems, see [6,8,9,10,11]. Researchers are primarily focused on understanding the division of periodic and quasi-periodic orbits, as well as the count and arrangement of limit cycles, sliding behaviour, grazing bifurcation, and chaos in discontinuous systems defined in two or many regions separated by single or multiple switching manifolds, see [12,13,14,15,16,17]. For example, in [12], a smooth system with nonsmooth perturbations was shown to develop a flat cone. The Poincaré map, the Melnikov-like function, and averaging theory have been developed as theoretically important tools for characterizing the bifurcation to a solution of the perturbed discontinuous system [18,19,20]. The Melnikov function, derived from the expansion of a Poincaré map, has been employed to examine the existence of crossing periodic orbits in discontinuous systems, as demonstrated in previous studies [9,21,22,23,24]. The Melnikov functions have recently been developed for nonlinear ship rolling systems that use delayed feedback controls and are excited by random waves [25,26].
The main contribution of this work is to develop a perturbation technique, which measures the effect of the nonsmooth perturbation on the PDB of the unperturbed system. To achieve this objective, the unperturbed system is supposed to be close to PDB, such that when the bifurcation parameter α passes through ˜α, the response changes from a period one to a period two limit cycle. We turn next to examine initially the necessary conditions for the existence of PDB in an unperturbed system by the so-called generalized Poincaré map. In this context, we derive the generalized determination of the Poincaré map associated to perturbed systems subjected to nonsmooth transitions. This map, computed here, is characterized by the composition of sub-maps that take into account the interaction with the discontinuity surface and its associated time intersection functions. We then introduce a function involving an unknown variable, where the zeros pinpoint the PDB that remains after the nonsmooth perturbation. Finally, we present a simple and novel discontinuous system that demonstrates how a nonsmooth perturbation causes a period doubling orbit to emerge from a flat periodic orbit. Another instance is given to demonstrate the formation of a perturbed period doubling orbit near an unperturbed one.
In a continuous nonlinear dynamical system, the slow variation of the bifurcation parameters can lead to instability, resulting in qualitative changes in the system's long-term behaviour. For instance, one of the most abundant complex and interesting behaviour is chaos. A system is classified as chaotic if it shows sensitivity to initial conditions and exhibits an infinite number of unstable periodic trajectories with varying periods [27,28]. One of the main ways chaos happens in parameter space is through a series of PDB, also known as a period doubling cascade. PDB occurs when the bifurcation parameter α passes through ˜α, an attracting periodic cycle becomes repelling and spawns an attracting orbit having twice the period, see Figure 1. As the bifurcation parameter is further modified, this period doubling may become unstable and give birth to a sequence of period doubling cycles.
Before stating our results for non-smooth differential models, we need to define and evaluate the bifurcation process for the period one to period two transition for such map. The goal of the following theorem is to outline a set of conditions that ensure the presence of PDB in a given map. Based on the results of a classical Period doubling bifurcation theorem, for one-dimensional smooth maps (i.e., d=1, see in [29, Theorem 3.5.1, p. 158]) and for the result for d>1, see [16], we present the following theorem to define the criteria for the existence of PDB in a continuous map.
Theorem 1. Assume that the mapping ˜P(ξ,α):Rd×R⟶Rd is smooth (i.e., it is uniformly continuous ˜P,∈Cr(Rd),r≥1). Then, ˜P has a PDB at (˜ξ,˜α) if ˜P satisfies the following conditions:
● The point (˜ξ,˜α) is a critical (fixed) point of ˜P, i.e., ˜P(˜ξ,˜α)=˜ξ.
● There exist U=[u1,u2]×X⊂Rd and θi:[u1,u2]→intX such that for α∈[˜α,α2] there holds
θ1(˜α)=θ2(˜α)=˜ξ(˜α),θ1(α)≠θ2(α),α≠˜α,˜P(θ1(α),α)=θ2(α),˜P(θ2(α),α)=θ1(α), |
where θ1(α),θ2(α), and ξ(α) are fixed points of ˜P(2)(ξ,α).
● The Jacobian matrix Dξ˜P(ξ,α)|(˜ξ,˜α) has eigenvalues λc(˜α)=−1 and λℓ(˜α),ℓ=1,2,...(d−1) with |λℓ(˜α)|≠1.
● Let us suppose ξ(α) represents the continuous curve of fixed points ˜P near ξ(˜α) and λℓ(α),and ℓ=1,2,...,(d−1) are the eigenvalues of the matrix Dξ˜P|(ξ(α)). Then, we have
μ=ddαλc(α)|˜α≠0. |
Therefore, there exists a smooth curve of fixed points, denoted as ξ(α), passing through ˜ξ at ˜α where the stability of the fixed point varies at ˜α (according to varying of μ).
Proof. Let us assume that the mapping or function Λ(ξ,α) is defined as:
Λ:c˜Σ+1→c˜Σ+1=(ξ,α)→˜P(ξ,α)−ξ, |
where c˜Σ+1 is an open subset of Rd, and we have Λ(˜ξ,˜α)=0 and ∂Λ∂ξ(˜ξ,˜α)≠0 associated to the eigenvalue λc(˜α)=−1. Given the second assumption that Dξ˜P has no eigenvalue equal to 1, we can use that the implicit function theorem to find a single solution of Λ(ξ,α)=0, near (˜ξ,˜α). This implies that there exists a curve of fixed points ξ(α) for Λ near ˜α such that ξ(˜α)=˜ξ.
Moreover, the Lyapunov-Schmidt reduction [30] can be used to obtain a bifurcation function F:R×R⊃dom(F)↔R, whose roots correspond to the fixed points and period two points of ˜P. For further fundamental tools on the bifurcation problem, we assume that ξ=(x,y)∈R×R(d−1), where x is a right eigenvector of Dξ˜P associated to the eigenvalue λc(α) (note that ˜x associated to λc(˜α)=−1). Our task here is to show that there exists a function y=y(x,α) defined on a set V=[x1,x2]×[α1,α2] with values in Y⊂R(d−1) such that y−πy(˜P(2)(x,y))=0 for all (x,y,α)∈V×Y, (˜P(2)(x,y) refers to the second iterate of the map ˜P(x,y)). Then, if DxF|(˜x,˜α)≠0, we can use the implicit function theorem to find the solution set of the bifurcation function F(x,α)=x−πx(˜P(2)(x,y(x,α)))=0,(x,α)∈V, and πx and πy are the coordinate projections onto the first and second factors of R×R(d−1). In addition, the bifurcation function when DxF|(˜x,˜α) has a non-trivial kernel, and then the situation can be simplified by applying the implicit function theorem to a new bifurcation function, that is G(x,α)=∫10DxF(t(x−ˉx(α))+ˉx(α),α)dt. This approach allows us to evaluate the solution of the bifurcation function as a regular curve and determine the fixed points of ˜P(2), i.e., the points of period two. Figure 1 is a bifurcation diagram showing the dynamical behavior of ˜P near the PDB point. It explains that the PDB occurs when a stable periodic orbit loses stability and undergoes bifurcation, resulting in the formation of a new periodic orbit with twice the period of the original orbit.
We consider the vector ξ∈Rn, with its time dependence described by the vector fields fi:Rn⟶Rn, i = 1, 2, and the switching function ℏ(ξ;ϵ) is a smooth function of both arguments, such that
˙ξ={f1(ξ;ϵ),ℏ(ξ;ϵ)<0,f2(ξ;ϵ),ℏ(ξ;ϵ)>0, | (3.1) |
where the phase space is divided into two domains separated by a hyperplane ˜Σ={ξ∈Rn∣ℏ(ξ;ϵ)=0}, h(ξ;ϵ) is a smooth function, and ϵ is a small parameter representing the size of the perturbation. In order to study (3.1), additional rules may be necessary to describe the interaction on the discontinuity surface ˜Σ, so this ensures that the flow of (3.1) is uniquely defined in the forward direction of time. Let ϱ(ξ;ϵ)=(nT(ξ;ϵ)f1(ξ;ϵ)).(nT(ξ;ϵ)f2(ξ;ϵ)), where the normal vector n(ξ;ϵ) perpendicular to the manifold ˜Σ is given as n(ξ;ϵ)=∇ℏ(ξ;ϵ)|∇ℏ(ξ;ϵ)|2, ‖n(ξ;ϵ)‖=1. Consequently, the discontinuity surface ˜Σ can be partitioned as follows:
(ⅰ) direct crossing region ˜Σc={ξ∈˜Σ∣ϱ(ξ;ϵ)>0}, more specifically ˜Σc=˜Σc−∪˜Σc+ such that ˜Σc±={ξ∈˜Σc∣±nT(ξ;ϵ)f1(ξ;ϵ)>0}.
(ⅱ) sliding region ˜Σs={ξ∈˜Σ∣ϱ(ξ;ϵ)≤0}, which in turn is divided into the attractive region ˜Σs−={ξ∈˜Σs∣nT(ξ;ϵ)f1(ξ;ϵ)>0}, and escaping region ˜Σs+={ξ∈˜Σs∣nT(ξ;ϵ)f2(ξ;ϵ)>0}. Further, ˜Σ0−={ξ∈˜Σ| nT(ξ;ϵ)f2(ξ,ϵ)=0},and˜Σ0+={ξ∈˜Σ|nT(ξ;ϵ)f1(ξ,ϵ)=0} are defined as the boundaries between sliding and crossing modes. Notice that ˜Σ is the disjoint union ˜Σc∪˜Σ0±∪˜Σs.
The sliding flows on ˜Σs are defined by the Filippov convex [31] combination as:
˙ξ=fs(ξ;ϵ)=nT(ξ;ϵ)f2(ξ;ϵ)⋅f1(ξ;ϵ)−nT(ξ;ϵ)f1(ξ;ϵ)⋅f2(ξ;ϵ)nT(ξ;ϵ)(f2(ξ;ϵ)−f1(ξ;ϵ)). | (3.2) |
Our objective is to examine the persistence of period doubling when (3.1) is perturbed within the category of all discontinuous vector fields with two regions separated by ˜Σ. A crossing period doubling of system (3.1) forms a closed path Γ, consisting of trajectories of fi,i=1,2 with matching directions, where Γ∩˜Σ=˜Σc.
The following is a general system that describes the behavior of a dynamical system close to PDB. Instead of working in the full space of the dynamical system, the dynamics are described in the transversal crossing trajectories branch.
Assume that it is possible to rewrite (3.1) as:
˙ξ=Xi(ξ)+ϵWi(ξ),(−1)iℏ(ξ;ϵ)<0,i=1,2, | (3.3) |
where X and W are uniformly continuous and bounded functions (X,W∈Cr(Rn∖˜Σ),r≥1). Let Ψ(k)(t(k)i(ξ,0),ξ;0) be the solutions of the unperturbed system (ϵ=0)
˙ξ=Xi(ξ),(−1)iℏ(ξ;0)<0. | (3.4) |
Then, the solution of (3.3) can be expressed as:
Ψ(k)i(τ(k)i(ξ;ϵ),ξ;ϵ)=Ψ(k)(τ(k)i(ξ,0),ξ;0)+ϵ∫τ(k)iτ(k)i−1Φ(k)i(τ(k)i)×(Φ(k)i)−1(s)Wi(Ψ(k)(τ(k)i(ξ,0),ξ;0),s)ds+o(ϵ2), | (3.5) |
where τ(k)i(ξ,0)=t(k)i(ξ), i=1,2,k=1,2 are the number of solution intersections with ˜Σ. The fundamental matrix solutions Φ(k)i(t(k)i) are given by solving the following linearized equations about (τ(k)i,ξ;ϵ)=(t(k)i,ξ0,0)
˙Φ(k)i(t(k))−∂Xi∂ξΦ(k)i(t(k)i)=0,Φ(k)i(0)=I, | (3.6) |
where I represents the identity matrix with the same order as the number of state variables.
Hypothesis 1. The unperturbed system (ϵ=0) is assumed to be close to a PDB, which means that the corresponding Poincaré map of (3.4) (this map is defined by the general solution (3.5) at ϵ=0) satisfies Theorem 1.
Without loss of generality, let us find a direction starting from the starting position ξ∈˜Σc−. Then, the solutions is denoted by Ψ(1)1(τ(1)1(ξ,ϵ),ξ;ϵ) which is Cr,r≥1. Moreover, assume that Ψ(1)1(τ(1)1(ξ;ϵ),ξ;ϵ) is within ˜Σc+ of the minimum return time τ(1)1(ξ;ϵ) at η=Ψ(1)1(τ(1)1(ξ;ϵ),ξ;ϵ)∈˜Σc+. Consequently, the flow given by Ψ(1)2(τ(1)2(η;ϵ),η;ϵ) is again within ˜Σc−. Thus, we define the first iterate of the Poincaré map as
˜P(1)(ξ;ϵ):˜Σc−→˜Σc−,˜P(1)(ξ;ϵ)=˜Π(1)2∘˜Π(1)1(ξ;ϵ), | (4.1) |
where
˜Π(1)1(ξ;ϵ):˜Σc−→˜Σc+,ξ(1)→Ψ(1)1(τ1(ξ;ϵ),ξ;ϵ)=η,˜Π(1)2(ξ;ϵ):˜Σc+→˜Σc−,ξ(2)→Ψ(1)2(τ1(ξ;ϵ),τ2(ξ;ϵ),η;ϵ)=ϑ. |
The perturbed system (3.1) has a period- one orbit if and only if ˜P(ˉξ(1),ϵ)=ˉξ(1) and τ(1)i exist. Since the period doubling orbit can be viewed as a fixed point of the second iteration of the Poincaré map, we have the mapping structure
˜P(2)(ξ;ϵ):˜Σc−→˜Σc−,˜P(2)(ξ;ϵ)=˜Π(2)2∘˜Π(2)1(ξ;ϵ), | (4.3) |
where
˜Π(2)1(ξ;ϵ):˜Σc−→˜Σc+,ξ(3)→Ψ(2)1(τ(1)2(ξ;ϵ),τ(2)1(ξ;ϵ),ϑ;ϵ)=χ,˜Π(2)2(ξ;ϵ):˜Σc+→˜Σc−,ξ(4)→Ψ(2)2(τ(2)1(ξ;ϵ),τ(2)2(ξ;ϵ),χ;ϵ)=ζ. |
It follows that the times of intersection exist for the first and second iteration of ˜P(2)(ξ;ϵ) by solving the following equations:
τ(k)i(ξ;ϵ):=inf{τ>0∣nT(ξ;ϵ)(˜Π(k)i(ξ;ϵ))=0},i,k=1,2. | (4.5) |
According to Theorem 1, we remark that ξ(1)≠ξ(2)≠ξ(3)≠ξ(4), and that ξ depends implicitly on ϵ. Because the absence of the transversality conditions at the intersection did not allow the trajectory of (3.3) to be a submanifold of the crossing area, having some sort of singular point. Therefore, the following argument shows that the transversality conditions are crucial to guarantee that ξ(l),l=1,2,3,4 and ˜Πki are not tangent to ˜Σ. In addition, these are sufficient conditions for an intersection to be stable after a perturbation.
Corollary 1. Let ˜Π(k)i:˜Σc→˜Σc be smooth maps, and ˜Σc a submanifold of ˜Σ. Then, ξ(l) and ˜Π(k)i are within the direct crossing set if the following conditions hold:
ϱ(ξ(l))>0,ϱ(˜Π(k)i)>0,i.e.,ξ(k)∈˜Σc,˜Π(k)i∈˜Σc, |
In general, because of the non-linearity of (4.1), (4.3), and (4.5), it is not possible to obtain an explicit expression for the sub-maps ˜Π(k)i. Thus, we are studying the characteristics of Poincaré sub-maps as they relate to the existence of period doubling orbit and the stability of the system (3.3).
Lemma 1. We assume that the transversality conditions that are given by the Corollary 1 hold. Then, for all ˉξ(i)∈˜Σc− and ˜Π(k)i(ˉξ(l),ϵ)∈˜Σc, i=1,2,k=1,2,l=1,...,4, the intersection function times τ(k)i(ξi;ϵ) are differentiable to ξ(l), so we get:
(a) Dξ˜Π(1)1(˜ξ(1);0)=(I−X1(˜ξ(1))nTnTX1(˜ξ(1)))Φ(1)1(˜t(1)1,˜ξ(1)),
(b) Dξ˜Π(1)2(˜ξ(2);0)=(I−X2(˜ξ(2))nTnTX2(˜ξ(2)))Φ(1)2(˜t(1)2,˜ξ(2))˜S(1)1×(I−X1(˜ξ(1))nTnTX1(˜ξ(1)))−1Dξ˜Π(1)1(˜ξ(1);0),
(c) Dξ˜Π(2)1(˜ξ(3);0)=(I−X1(˜ξ(3))nTnTX1(˜ξ(3)))Φ(2)1(˜t(2)1,˜ξ(3))˜S(2)1×(I−X2(˜ξ(2))nTnTX2(˜ξ(2)))−1Dξ˜Π(1)2(˜ξ(2);0),
(d) Dξ˜Π(2)2(˜ξ(4);0)=(I−X2(˜ξ(4))nTnTX2(˜ξ(4)))Φ(2)2(˜t(2)2,˜ξ(4))˜S(2)2×(I−X1(˜ξ(3))nTnTX1(˜ξ(3)))−1Dξ˜Π(2)1(˜ξ(3);0).
Proof. Consider the partial derivatives of ˜Π(k)i(ξ;ϵ) with respect to ξ at the point (τ(k)i,ξ(l);ϵ)=(˜ti(k),˜ξ(l),0), where Dξ=∂∂ξ:
Dξ˜Π(1)1(˜ξ(1);0)=Φ(1)1(˜t(1)1,˜ξ(1))+X1(˜ξ(1))Dξτ(1)1(˜ξ(1);0)Dξ˜Π(1)2(˜ξ(2);0)=Dτ(1)1Ψ(1)2(˜t(1)2,˜ξ(2);0)Dξτ(1)1(˜ξ(1);0)+X2(˜ξ(2))Dξτ(1)2(˜ξ(2);0)+Φ(1)2(˜t(1)2,˜ξ(2))Dξ˜Π(1)1(˜ξ(1);0),Dξ˜Π(2)1(ˉξ(3);0)=Dτ(1)2Ψ(2)1(˜t(2)1,˜ξ(3);0)Dξτ(1)2(˜ξ(2);0)+X1(˜ξ(3))Dξτ(2)1(˜ξ(3);0)+Φ(2)1(˜t(2)1,˜ξ(3))Dξ˜Π(1)2(˜ξ(2);0),Dξ˜Π(2)2(ˉξ(4);0)=Dτ(2)1Ψ(2)2(˜t(2)2,˜ξ(4);0)Dξτ(2)1(˜ξ(3);0)+X2(˜ξ(4))Dξτ(2)2(˜ξ(4);0)+Φ(2)2(˜t(2)2,˜ξ(4))Dξ˜Π(2)1(ˉξ(3);0). | (4.6) |
In order to find the explicit formulas of Dξ˜Π(k)i, we have to compute the terms Dξτ(k)i. The partial derivatives of τ(k)i are obtained by noting that nTDξ˜Π(k)i(˜ξ(l);0)=0. Therefore, from nTDξ˜Π(1)1(˜ξ(1);0)=0, we obtain Dξτ(1)1(˜ξ(1);0)=−nTΦ1(~t1(1),˜ξ(1))nTX1(˜ξ(1)), and (a) holds.
At the intersection points, we have Dτ(1)i(˜ξ(i+1);0)Ψ(i)3−i=−Φ(i)3−iX3−i. Then, the result of the equation nTDξ˜Π(1)2=0 implies
nT(Φ(1)2(˜t(1)2,˜ξ(2))(I+(X2(˜ξ(1))−X1(˜ξ(1)))nTnTX1(˜ξ(1)))⏟˜S(1)1Φ(1)1(~t1(1),˜ξ(1))+X2(˜ξ(2))Dξτ(1)2(˜ξ(2);0))=0. |
Then, we obtain Dξτ(1)2(˜ξ(2);0)=−nTΦ(1)2(~t2(1),˜ξ2)˜S(1)1Φ(1)1(~t1(1),˜ξ1)nTX2(˜ξ2), and the direct substitution into the second equation of (4.6) leads to statement (b).
In the same way, we are going to solve nTDξ˜Π(2)i(˜ξ(2+i);0)=0 for Dξτ(i)2, which gives: Dξτ(2)1=−nTΦ(2)1(˜t(2)1,˜ξ3)˜S(1)2Φ(1)2(˜t(1)2,˜ξ2)˜S(1)1Φ(1)1(˜t(1)1,˜ξ1)nTX1(˜ξ3), Dξτ(2)2=−nTΦ(2)2(~t2(2),˜ξ(4))˜S(2)1Φ(2)1(~t1(2),˜ξ3)˜S(1)2Φ(1)2(~t2(1),˜ξ2)˜S(1)1Φ1(~t1(1),˜ξ1)nTX2(˜ξ4).
Substituting Dξτ(2)i in the third and fourth formulas of (4.6), we will ultimately obtain formulas (c) and (d), respectively.
The following theorem illustrates that, by linearizing the Poincaré map, we can carry out explicit analytical calculations.
Theorem 2. Assume that ξ∈˜Σc− and ˜P(m):˜Σc−→˜Σc−, such that ˜P(m)(0;0)=0 and ˜P(m)(˜ξ;0)=˜ξ. Then, the linearization of the Poincaré map is given by
Dξ˜P(m)=m∏v=1˜S(m+1−v)2Φ(m+1−v)2˜S(m+1−v)1Φ(m+1−v)1. |
Further, ˜P(m)(˜ξ;0)X1(˜ξ)=X1(˜ξ) and attractivity of the period-m orbit is determined by the remaining (n-2) eigenvalues of Dξ˜P(m).
Proof. Assume that ˜ξ∈˜Σc−. The natural condition for period doubling orbit to exist in an unperturbed system is Ψ(1)1(0,˜ξ;0)=Ψ(1)1(T,˜ξ;0), or in a compact mapping as
˜P(2)(˜ξ;0)=˜Π(2)2∘˜Π(2)1∘˜Π(1)2∘˜Π(1)1(˜ξ;0)=˜ξ, |
where T=∑2k=1∑2i=1τ(k)i(˜ξ;0)=∑2k=1∑2i=1t(k)i(ξ). We assume that the second iteration of the Poincaré map is provided as:
˜P(2)(ξ;ϵ)=˜Π(2)2(ξ;ϵ)+∫τ(1)1τ(2)2(X1(ξ,s)+ϵW1(ξ,s;ϵ))ds. | (4.7) |
Differentiating (4.7) with respect to ξ, we obtain
Dξ˜P(2)=Dξ˜Π(2)2(ξ;ϵ)−(X1(ξ)+ϵW1(ξ;ϵ))Dξτ(2)2(ξ;ϵ)+∫τ(1)1˜τ(2)2(X1(ξ,s)+ϵW1(ξ,s;ϵ))ds. |
The above derivative is estimated at (τ,ξ,ϵ)=(˜t,˜ξ,0) as
Dξ˜P(2)=Dξ˜Π(2)2(ξ;0)−X1(ξ)Dξτ(2)2(ξ;0)+∫τ(1)1˜τ(2)2X1(ξ,s)ds. |
Using the results of Lemma 1 and the value of the vector function Dξτ(2)2(ξ;0), we obtain
Dξ˜P(2)=2∏v=1˜S(3−v)2Φ(3−v)2˜S(3−v)1Φ(3−v)1+∫τ(1)1˜τ(2)2X1(ξ,s)ds. |
At the second iteration of the Poincaré map, substituting τ(1)1 into ˜t(2)2,
Dξ˜P(2)=2∏v=1˜S(3−v)2Φ(3−v)2˜S(3−v)1Φ(3−v)1. |
In this situation, when the unperturbed system has period-m orbits, it is simply a matter of multiplying the matrices Dξ˜P(k)i in the correct order. Then, we obtain
Dξ˜P(m)=m∏v=1˜S(m+1−v)2Φ(m+1−v)2˜S(m+1−v)1Φ(m+1−v)1. |
Remark 1 It should be noted that, when m=1, a similar result has obtained in [9,19].
Lemma 2. Suppose that Γ⊂˜Σc− is a period doubling orbit of the unperturbed system generated by (3.3), and the transversality conditions that are given by Corollary 1 hold. Then, for all ˉξ(i)∈˜Σc− and ˜Π(k)i(ˉξ(l),ϵ)∈˜Σc, i=1,2,k=1,2,l=1,...,4, and the intersection times functions τ(k)i(ξi;ϵ) are differentiable to ϵ, so we get:
(a) Dϵ˜Π(1)1(˜ξ(1);0)=(I−X1(˜ξ(1))nTnTX1(˜ξ(1)))DϵΨ(1)1(˜t(1)1,˜ξ(1)),
(b) Dϵ˜Π(1)2(˜ξ(2);0)=(I−X2(˜ξ(2))nTnTX2(˜ξ(2)))(Φ(1)2(˜t(1)2,˜ξ(2))˜S(1)1×(I−X1(˜ξ(1))nTnTX1(˜ξ(1)))−1Dϵ˜Π(1)1(˜ξ(1);0)+DϵΨ(1)2(˜t(1)2,˜ξ(2))),
(c) Dϵ˜Π(2)1(˜ξ(3);0)=(I−X1(˜ξ(3))nTnTX1(˜ξ(3)))(Φ(2)1(˜t(2)1,˜ξ(3))˜S(1)2×(I−X2(˜ξ(2))nTnTX2(˜ξ(2)))−1Dϵ˜Π(1)2(˜ξ(2);0)+DϵΨ(2)1(˜t(2)1,˜ξ(3))),
(d) Dϵ˜Π(2)2(˜ξ(4);0)=(I−X2(˜ξ(4))nTnTX2(˜ξ(4)))(Φ(2)2(˜t(2)2,˜ξ(4))˜S(2)2×(I−X1(˜ξ(3))nTnTX1(˜ξ(3)))−1Dϵ˜Π(2)1(˜ξ(3);0)+DϵΨ(2)2(˜t(2)2,˜ξ(4))).
Proof. Consider the partial derivatives of ˜Π(k)i(ξ;ϵ) with respect to ϵ at the point (τ(k)i,ξ(l);ϵ)=(˜ti(k),˜ξ(l),0),
Dϵ˜Π(1)1(˜ξ(1);0)=DϵΨ(1)1(˜t(1)1,˜ξ(1))+X1(˜ξ(1))Dϵτ(1)1(˜ξ(1);0),Dϵ˜Π(1)2(˜ξ(2);0)=Dτ(1)1Ψ(1)2(˜t(1)2,˜ξ(2);0)Dϵτ(1)1(˜ξ(1);0)+X2(˜ξ(2))Dϵτ(1)2(˜ξ(2);0)+Φ(1)2(˜t(1)2,˜ξ(2))Dϵ˜Π(1)1(˜ξ(1);0)+DϵΨ(1)2(˜t(1)2,˜ξ(2)),Dϵ˜Π(2)1(ˉξ(3);0)=Dτ(1)2Ψ(2)1(˜t(2)1,˜ξ(3);0)Dϵτ(1)2(˜ξ(2);0)+X1(˜ξ(3))Dϵτ(2)1(˜ξ(3);0)+Φ(2)1(˜t(2)1,˜ξ(3))Dϵ˜Π(1)2(˜ξ(2);0)+DϵΨ(2)1(˜t(2)1,˜ξ(3)),Dϵ˜Π(2)2(ˉξ(4);0)=Dτ(2)1Ψ(2)2(˜t(2)2,˜ξ(4);0)Dϵτ(2)1(˜ξ(3);0)+X2(˜ξ(4))Dϵτ(2)2(˜ξ(4);0)+Φ(2)2(˜t(2)2,˜ξ(4))Dϵ˜Π(2)1(ˉξ(3);0)+DϵΨ(2)2(˜t(2)2,˜ξ(4)). | (4.8) |
To complete the computation of the above sub-maps, we need to calculate the terms Dϵτ(k)i(˜ξ(l);0) and DϵΨ(k)i(˜t(k)i,˜ξ(l)). The derivative of (3.3) for ϵ is given as:
˙ξϵ=DξXiξϵ+Wi(ξ;0). | (4.9) |
Note that the recurrence Eq (4.9) describes initial value problems such as ξϵ(0)=0. The general solution (3.5) is therefore equivalent to the following recurrence:
DϵΨ(k)i=∫τ(k)iτ(k)i−1Φ(k)i(τi)(Φ(k)i)−1(s)Wi(Ψ(k)i(τ(k)i(ξ,0),ξ,0),s)ds. | (4.10) |
In addition, the partial derivatives of Dϵτ(k)i(˜ξ(l);0) are determined by solving the equations nTDϵ˜Π(k)i(˜ξ(l);0)=0. So, we get the following at the intersection point ξ(l)∈˜Σc:
Dϵτ(1)1=−nTDϵΨ(1)1nTX1(˜ξ(1)),Dϵτ(1)2=−nT(Φ(1)2(~t2(1),˜ξ(2))˜S(1)1DϵΨ(1)1+DϵΨ(1)2)nTX2(˜ξ(2)),Dϵτ(2)1=−nTnTX1(˜ξ(3))(Φ(2)1(˜t(2)1,˜ξ(3))˜S(1)2Φ(1)2(˜t(1)2,˜ξ(2))˜S(1)1DϵΨ(1)1+Φ(2)1(˜t(2)1,˜ξ(3))˜S(1)2DϵΨ(1)2+DϵΨ(2)1),Dϵτ(2)2=−nTnTX2(˜ξ(4))(DϵΨ(2)2+Φ(2)2(~t2(2),˜ξ(4))˜S(2)1DϵΨ(2)1+Φ(2)2(~t2(2),˜ξ(4))˜S(2)1Φ(2)1(~t1(2),˜ξ(3))˜S(1)2Φ(1)2(~t2(1),˜ξ(2))˜S(1)1DϵΨ(1)1+Φ(2)2(~t2(2),˜ξ(4))˜S(2)1Φ(2)1(~t1(2),˜ξ(3))˜S(1)2DϵΨ(1)2). | (4.11) |
Through substituting (4.10) and (4.11) into (4.8), we get the equalities (a)–(d).
The following theorem presents the derivation of the Melnikov function associated with perturbation limit cycles and PDB of a differential system (3.3).
Theorem 3. Assume that ξ∈˜Σc− and ˜P(i):˜Σc−→˜Σc−, such that ˜P(1)(0;0)=0 and ˜P(2)(˜ξ;0)=˜ξ. Then,
Dϵ˜P(1)=˜S(1)22∑k=1Φ(2−k)2˜S(2−k)1∂Ψ(1)k∂ϵDϵ˜P(2)=(2∏v=1˜S(3−v)2Φ(3−v)2˜S(3−v)1Φ2(2−v)1)∂Ψ(1)2∂ϵ+(2∏v=1˜S2(2−v)2Φ2(2−v)2˜S2(2−v)1Φ2(2−v)1˜S(2−v)1)∂Ψ(1)1∂ϵ+(2∏v=1˜S2(2−v)2Φ2(2−v)2˜S2(2−v)1)∂Ψ(2)1∂ϵ+(2∏v=1˜S2(2−v)2)∂Ψ(2)2∂ϵ. |
Proof. Differentiating (4.7) with respect to ϵ at (τ,ξ,ϵ)=(˜t,˜ξ,0)), we obtain
Dϵ˜P(2)=Dϵ˜Π(2)2(˜ξ(4);0)−X1(˜ξ(4))Dϵτ(2)2(˜ξ(4);0). |
Using the results of Lemma 2 and the value of the vector function Dξτ(2)2(ξ;0) from (4.11), the proof of the statement of this theorem is straightforward.
It should be noted that a similar result for the expansion of the Melnikov function Dϵ˜P(1) that detects the existence of limit cycles is obtained in [9,19]. We present the following corollary, which is based on the concepts of orthogonal projection and related Melnikov function properties discussed in [19].
Corollary 2. Assume that ˜ξ∈˜Σc− and eT1˜ξ=0. Then, the orthogonal projection along one-dimensional space [X1(˜ξ)] is given as ψX1V=⟨V,X1(˜ξ)⟩X1(˜ξ)⟨X1(˜ξ),X1(˜ξ)⟩, and the orthogonal projection complement is given by ψX⊥1V=I−⟨V,X1(˜ξ)⟩X1(˜ξ)⟨X1(˜ξ),X1(˜ξ)⟩. Therefore, the linearized Poincaré map of the original system is given by
Dξ˜˜P(m)=(I−ψX1)Dξ˜P(m),Dϵ˜˜P(i)=(I−ψX1)Dϵ˜P(i). | (4.12) |
In the following section, we present a novel discontinuous systems that highlights how the nonsmooth perturbation is responsible for the period doubling orbit that emerges from an existing flat periodic orbit (or near an unperturbed one).
Because it is difficult to obtain an explicit analytical expression of the generalized Poincaré map, the theoretical results are numerically validated using path-following techniques for discontinuous dynamical systems. There are numerically sound techniques for approximating the solutions of discontinuous dynamical systems. For examples, see [32,33,34].
Example 1. We consider a simple discontinuous system with a switching plane ˜Σ={ξ∈R3∣eT1ξ=0}, which is written as:
˙ξ=Aiξ+ϵWi(ξ),(−1)ieT1ξ<0,i=1,2, | (5.1) |
where Ai are 3×3 matrices defined as:
Ai=(0ω0−ω0000μ1),W1(ξ)=(00y2),W2(ξ)=(0μ2z2μ3y2),ω,μv∈R,v=1,2,3, |
such that Wi(0)=0, and the nonlinear perturbations Wi(ξ)=∘(∥ξ∥) are Ck-maps k≥1 defined on the whole phase space R3.
We note that the origin of coordinates is still a unique equilibrium point for both systems, and the eigenvalues of the linearized systems ˙ξ=Aiξ are given as λ1=μ1,λ2,3=±ˉiω,ˉi2=−1. Further, both linear smooth systems (ϵ=0) possess an invariant plane z=0 with constant return times τ±=π. Therefore, if ϵ=0, only one family of flat periodic orbits is generated by the eigenvector: ξ=(0,ˉy,0),ˉy∈R.
For ϵ≠0, μ1=0.15, and μ3=−0.6, the perturbed system (5.1) has a period doubling orbit that emerges from an existing flat periodic orbit with approximately twice the period of the flat orbit. Here, we have Figure 2 with ω=μ2=1 and ϵ=0.15, Figure 3 with ω=μ2=1 and ϵ=0.2, and Figure 4 with ω=μ2=−1 ϵ=0.5.
The following example illustrates the emergence of a perturbed period doubling orbit near an unperturbed one.
Example 2. Consider the discontinuous system
˙ξ=Xi(ξ)+ϵWi(ξ),(−1)ieT1ξ<0, | (5.2) |
where
Xi(ξ)=(yz−az−(−1)iy2−x),Wi(ξ)=(00y(x−(−1)iz)), |
a>0, and the phase space is divided into two domains separated by a hyperplane ˜Σ={ξ∈R3∣ℏ(ξ;ϵ)=x=0}. Further, the perturbed system (5.2) has only the crossing mode and cannot exhibit a sliding mode due to ˜Σs=∅. By numerical computation, it is straightforward to check that the unperturbed system of (5.2) has a PDB at a=1.7, see Figure 5 at ϵ=0. By setting the perturbation value to ϵ=0.1, numerical computations show that the perturbed system (5.2) has a period doubling orbit near the unperturbed period doubling orbit, see Figure 5 at ϵ=0.1.
Practical applications in the real world motivate the development of novel mathematical techniques for analyzing the dynamics of nonsmooth systems. Existing smooth methods, for example, lack a mechanism that explains how switching manifolds generate period-doubling orbits. This paper describes a novel application of the Melnikov method for discontinuous differential systems separated by a hypersurface, with the goal of determining the effect of a nonsmooth perturbation on the period doubling bifurcation of an unperturbed system. The generalized Poincaré map for perturbed systems undergoing nonsmooth transitions has been derived, assuming transversal intersections with discontinuity-switching boundaries and the unperturbed system having a period-doubling orbit. The Melnikov function is designed for systems that have discontinuities and acts as an extension of the regular Melnikov function used in smooth scenarios. Two examples are provided to numerically validate the theoretical results, showing how non-smooth perturbations can cause a period doubling orbit to emerge from an existing flat periodic orbit, as well as how a new period doubling orbit emerges close to the original.
Hany A. Hosham: Conceptualization, formal analysis, investigation, software, writing-original draft; Alaa A. Alzulaibani: Data curation, funding acquisition, project administration, writing-review & editing; Tarek Sellami: Supervision, investigation, writing-review & editing; Khaled Sioud: Methodology, supervision; Thoraya N. Alharthi: Visualisation, formal analysis, methodology, writing-original draft. 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 article.
The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research work through the project number 445-9-734.
The authors declare no conflict of interest in this paper.
[1] |
H. Svensmark, M. R. Samuelsen, Perturbed period-doubling bifurcation. Ⅰ. Theory, Phys. Rev. B, 41 (1990), 4181–4188. https://doi.org/10.1103/PhysRevB.41.4181 doi: 10.1103/PhysRevB.41.4181
![]() |
[2] |
A. J. Sojahrood, R. Earl, M. C. Kolios, R. Karshafian, Investigation of the 1/2 order subharmonic emissions of the period-2 oscillations of an ultrasonically excited bubble, Phys. Lett. A, 384 (2020), 126446. https://doi.org/10.1016/j.physleta.2020.126446 doi: 10.1016/j.physleta.2020.126446
![]() |
[3] |
C. Athanasouli, K. Kalmbach, V. Booth, C. G. D. Behn, Nrem-rem alternation complicates transitions from napping to non-napping behavior in a three-state model of sleep-wake regulation, Math. Biosci., 355 (2023), 108929. https://doi.org/10.1016/j.mbs.2022.108929 doi: 10.1016/j.mbs.2022.108929
![]() |
[4] |
N. F. Tehrani, M. R. Razvan, Bifurcation structure of two coupled FHN neurons with delay, Math. Biosci., 270 (2015), 41–56. https://doi.org/10.1016/j.mbs.2015.09.008 doi: 10.1016/j.mbs.2015.09.008
![]() |
[5] |
L. Gyllingberg, D. J. T. Sumpter, Å. Brännström, Finding analytical approximations for discrete, stochastic, individual-based models of ecology, Math. Biosci., 365 (2023), 109084. https://doi.org/10.1016/j.mbs.2023.109084 doi: 10.1016/j.mbs.2023.109084
![]() |
[6] | M. D. Bernardo, C. J. Budd, A. R. Champneys, P. Kowalczyk, Piecewise-smooth dynamical systems theory and applications, Springer Science & Business Media, 163 (2018). |
[7] |
T. Küpper, H. A. Hosham, K. Dudtschenko, The dynamics of bells as impacting system, Proc. I. Mech. Eng. Part, 225 (2011), 2436–2443. https://doi.org/10.1177/0954406211413855 doi: 10.1177/0954406211413855
![]() |
[8] |
J. Awrejcewicz, M. Fečkan, P. Olejnik, Bifurcations of planar sliding homoclinics, Math. Probl. Eng., 2006 (2006), 1–13. https://doi.org/10.1155/MPE/2006/85349 doi: 10.1155/MPE/2006/85349
![]() |
[9] |
H. A. Hosham, Bifurcation of periodic orbits in discontinuous systems, Nonlinear Dynam., 87 (2017), 135–148. https://doi.org/10.1007/s11071-016-3031-7 doi: 10.1007/s11071-016-3031-7
![]() |
[10] |
H. A. Hosham, T. N. Alharthi, Bifurcation and chaos in simple discontinuous systems separated by a hypersurface, AIMS Math., 9 (2024), 17025–17038. https://doi.org/10.3934/math.2024826 doi: 10.3934/math.2024826
![]() |
[11] |
M. R. Jeffrey, T. I. Seidman, M. A. Teixeira, V. I. Utkin, Into higher dimensions for nonsmooth dynamical systems, Physica D, 434 (2022), 133222. https://doi.org/10.1016/j.physd.2022.133222 doi: 10.1016/j.physd.2022.133222
![]() |
[12] |
T. Küpper, H. A. Hosham, Reduction to invariant cones for non-smooth systems, Math. Comput. Simulat., 81 (2011), 980–995. https://doi.org/10.1016/j.matcom.2010.10.004 doi: 10.1016/j.matcom.2010.10.004
![]() |
[13] |
D. Weiss, T. Küpper, H. A. Hosham, Invariant manifolds for nonsmooth systems, Physica D, 241 (2012), 1895–1902. https://doi.org/10.1016/j.physd.2011.07.012 doi: 10.1016/j.physd.2011.07.012
![]() |
[14] |
D. Weiss, T. Küpper, H. A. Hosham, Invariant manifolds for nonsmooth systems with sliding mode, Math. Comput. Simulat., 110 (2015), 15–32. https://doi.org/10.1016/j.matcom.2014.02.004 doi: 10.1016/j.matcom.2014.02.004
![]() |
[15] |
H. A. Hosham, Bifurcation of limit cycles in piecewise-smooth systems with intersecting discontinuity surfaces, Nonlinear Dynam., 99 (2020), 2049–2063. https://doi.org/10.1007/s11071-019-05400-z doi: 10.1007/s11071-019-05400-z
![]() |
[16] | H. A. Hosham, Nonlinear behavior of a novel switching jerk system, Int. J. Bifurc. Chaos, 2020. https://doi.org/10.1142/S0218127420502028 |
[17] |
F. Luo, Z. D. Du, Complicated periodic cascades arising from double grazing bifurcations in an impact oscillator with two rigid constraints, Nonlinear Dynam., 111 (2023), 13829–13852. https://doi.org/10.1007/s11071-023-08600-w doi: 10.1007/s11071-023-08600-w
![]() |
[18] | Z. Fan, Z. D. Du, Bifurcation of periodic orbits crossing switching manifolds multiple times in planar piecewise smooth systems, Int. J. Bifurcat. Chaos, 29 (2019). https://doi.org/10.1142/S0218127419501608 |
[19] | M. Feckan, M. Pospíšil, Poincaré-Andronov-Melnikov analysis for non-smooth systems, Oxford: Academic Press, 2016. |
[20] |
J. Llibre, D. D. Novaes, C. A. B. Rodrigues, Bifurcations from families of periodic solutions in piecewise differential systems, Physica D, 404 (2020), 132342. https://doi.org/10.1016/j.physd.2020.132342 doi: 10.1016/j.physd.2020.132342
![]() |
[21] |
X. Guo, R. Tian, Q. Xue, X. Zhang, Sub-harmonic Melnikov function for a high-dimensional non-smooth coupled system, Chaos Soliton. Fract., 164 (2022), 112629. https://doi.org/10.1016/j.chaos.2022.112629 doi: 10.1016/j.chaos.2022.112629
![]() |
[22] |
J. L. R. Bastos, C. A. Buzzi, J. Llibre, D. D. Novaes, Melnikov analysis in nonsmooth differential systems with nonlinear switching manifold, J. Differ. Equations, 267 (2019), 3748–3767. https://doi.org/10.1016/j.jde.2019.04.019 doi: 10.1016/j.jde.2019.04.019
![]() |
[23] |
K. S. Andrade, O. A. R. Cespedes, D. R. Cruz, D. D. Novaes, Higher order Melnikov analysis for planar piecewise linear vector fields with nonlinear switching curve, J. Differ. Equations, 287 (2021), 1–36. https://doi.org/10.1016/j.jde.2021.03.039 doi: 10.1016/j.jde.2021.03.039
![]() |
[24] | Y. Li, Z. Du, Applying Battelli-Fečkan's method to transversal heteroclinic bifurcation in piecewise smooth systems, Discrete Cont. Dyn.-S., 24 (2019), 6025–6052. |
[25] |
M. Wang, Z. Wei, J. Wang, X. Yu, T. Kapitaniak, Stochastic bifurcation and chaos study for nonlinear ship rolling motion with random excitation and delayed feedback controls, Physica D, 462 (2024), 134147. https://doi.org/10.1016/j.physd.2024.134147 doi: 10.1016/j.physd.2024.134147
![]() |
[26] |
Z. Wei, Y. Li, T. Kapitaniak, W. Zhang, Analysis of chaos and capsizing of a class of nonlinear ship rolling systems under excitation of random waves, Chaos Int. J. Nonlinear Sci., 34 (2024), 043106. https://doi.org/10.1063/5.0187362 doi: 10.1063/5.0187362
![]() |
[27] | S. H. Strogatz, Nonlinear dynamics and chaos: With applications to physics, biology, chemistry, and engineering, 3 Eds., CRC press, 2024. |
[28] | R. L. Devaney, An introduction to chaotic dynamical systems, 2 Eds., CRC press, 2018. https://doi.org/10.4324/9780429502309 |
[29] | J. Guckenheimer, P. Holmes, Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, New York: Springer, 1984. https://doi.org/10.1007/978-1-4612-1140-2 |
[30] |
C. Chicone, Lyapunov-schmidt reduction and melnikov integrals for bifurcation of periodic solutions in coupled oscillators, J. Differ. Equations, 112 (1994), 407–447. https://doi.org/10.1006/jdeq.1994.1110 doi: 10.1006/jdeq.1994.1110
![]() |
[31] |
P. Kowalczyk, M. D. Bernardo, Two-parameter degenerate sliding bifurcations in Filippov systems, Physica D, 204 (2005), 204–229. https://doi.org/10.1016/j.physd.2005.04.013 doi: 10.1016/j.physd.2005.04.013
![]() |
[32] | V. Acary, B. Brogliato, Numerical methods for nonsmooth dynamical systems: Applications in mechanics and electronics, Springer Science & Business Media, 2008. https://doi.org/10.1007/978-3-540-75392-6 |
[33] |
L. Dieci, L. Lopez, Numerical solution of discontinuous differential systems: Approaching the discontinuity surface from one side, Appl. Numer. Math., 67 (2013), 98–110. https://doi.org/10.1016/j.apnum.2011.08.010 doi: 10.1016/j.apnum.2011.08.010
![]() |
[34] |
N. Guglielmi, E. Hairer, An efficient algorithm for solving piecewise-smooth dynamical systems, Numer. Algorithms, 89 (2022), 1311–1334. https://doi.org/10.1007/s11075-021-01154-1 doi: 10.1007/s11075-021-01154-1
![]() |