Stationary response of stochastic viscoelastic system with the right unilateral nonzero offset barrier impacts
Wang Deli1, Xu Wei1, †, Gu Xudong2
Department of Applied Mathematics, Northwestern Polytechnical University, Xi’an 710072, China
Department of Engineering Mechanics, Northwestern Polytechnical University, Xi’an 710072, China

 

† Corresponding author. E-mail: weixu@nwpu.edu.cn

Project supported by the National Natural Science Foundation of China (Grant Nos. 11872305 and 11872307) and the Excellent Doctorate Cultivating Foundation of Northwestern Polytechnical University, China.

Abstract

The stationary response of viscoelastic dynamical system with the right unilateral nonzero offset barrier impacts subjected to stochastic excitations is investigated. First, the viscoelastic force is approximately treated as equivalent terms associated with effects. Then, the free vibro-impact (VI) system is absorbed to describe the periodic motion without impacts and quasi-periodic motion with impacts based upon the level of system energy. The stochastic averaging of energy envelope (SAEE) is adopted to seek the stationary probability density functions (PDFs). The detailed theoretical results for Van der Pol viscoelastic VI system with the right unilateral nonzero offset barrier are solved to demonstrate the important effects of the viscoelastic damping and nonzero rigid barrier impacts condition. Monte Carlo (MC) simulation is also performed to verify the reliability of the suggested approach. The stochastic P-bifurcation caused by certain system parameters is further explored. The variation of elastic modulus from negative to zero and then to positive witnesses the evolution process of stochastic P-bifurcation. From the vicinity of the common value to a wider range, the relaxation time induces the stochastic P-bifurcation in the two interval schemes.

1. Introduction

Viscoelastic materials, such as polymer (viscoelastic anti-corrosion adhesive tape), synthetic material (synthetic rubber and geomembrane), and metallurgical material (brake pad and loose-hole sheet), with the advantages of anti-corrosion,[1] shock absorption, and noise reduction,[2,3] are favored in practical applications from life and production to construction engineering, aerospace, medical science, and other fields. Thus, viscoelastic property has attracted a lot of attention in scientific research. For viscoelastic materials, the stress depends on the history of strain, which is different from the classical elasticity.[4] It could be designed as a damping layer where the characteristic depends on frequency and temperature.[5] It has been observed that the strain of a viscoelastic material under constant stress increases with time in creep tests, while the stress decreases with time under constant strain in relaxation tests.[6] Based on such consideration, a generalized Maxwell model[7,8] has been developed to describe the relaxation function.

In the previous literature, the nonlinear dynamic study of viscoelastic systems has mainly focused on the deterministic case.[911] Due to the widespread existence of random perturbations in practical applications,[1214] it is quite significant to investigate the behavior of viscoelastic systems under stochastic excitations. Potapov[15] analyzed the stability of stochastic viscoelastic systems by the method of stochastic averaging. Cai and Zhu[16] improved the stochastic averaging method based on the viscoelastic system. Huang et al.[17] studied the stability of a single-degree-of-freedom (SDOF) linear viscoelastic system by using both first-order and second-order stochastic averaging methods. Mario and collaborators[18] considered stationary and non-stationary stochastic response of linear fractional viscoelastic systems. Christian and Poloskov[19] provided time-domain formulation in computational dynamics for linear viscoelastic media with model uncertainties and stochastic excitation.

In addition, a number of non-smooth factors[20] exist in engineering applications; for instance, the interactions of a mechanical or structural system with a rigid obstacle will result in a non-smooth effect, which is modeled as impact. The interaction of the system with obstacles can be either elastic or inelastic.[21] Thus, the dynamics of vibro-impact (VI) systems has been further studied.[2228] Zhuralev has introduced a non-smooth transformation to study the VI system with rigid barriers,[29] and Wu and Zhu extended this method to multiple-degrees-of-freedom (MDOF) VI systems under Poisson white noise excitations;[30] some further developments of this method have been made subsequently.[3133] Chattering bifurcations were considered in a Duffing unilateral VI system.[34] The finite element method and a measure based on the Shannon entropy[3537] were described in a noisy VI Duffing–Van der Pol oscillator. Resonance response of a nonlinear VI system to a narrow-band random parametric excitation are analyzed in Ref. [38]. In addition, the studies on stochastic bifurcation in Refs. [39]–[44] are referenced. Most of the previous studies have concentrated on the responses of linear VI systems with zero rigid barrier or nonlinear smooth systems, and only a few considered the responses of nonlinear nonzero offset barrier impacts oscillator coupled with the viscoelastic force.

This paper mainly focuses on the stationary response considerations of the stochastic viscoelastic system with right nonzero offset barrier impacts. The structure of the paper is as follows. The original system and viscoelastic damping treatment are described in Section 2. In Section 3, the free VI system is analyzed, and then the approximate stationary probability density functions (PDFs) are derived by the stochastic averaging approach of energy envelope. An example of Van der Pol viscoelastic VI oscillator with the right nonzero rigid barrier is carried out and the analytical solutions observed from the suggested method are well tested by the results from the Monte Carlo (MC) simulation in Section 4.1, and the stochastic P-bifurcation caused by certain parameters is explored in Section 4.2. The conclusion is summarized in the last section.

2. Problem statement
2.1. Original system description

An SDOF lightly damped viscoelastic system with right unilateral nonzero offset barrier impacts behavior under random excitations is investigated, which is schematically shown in Fig. 1. The equation of the system is of the form

Fig. 1. Schematic model of the viscoelastic oscillator with right unilateral nonzero offset barrier impacts.

In Eq. (1a), ε is a small positive parameter, and the dot indicates differentiation with respect to time; represents lightly linear or nonlinear damping force; r(x) denotes the nonlinear or linear restoring force as an odd function of x; ηm (t) represents Gaussian white noises in the sense of Stratonovich, and possesses correlation functions E[ηn(t)ηk (t + τ0)] = 2Dnk δ (τ0); and (mN+) describes amplitudes of weakly random perturbations. Equation (1b) is the impact condition for the inelastic instantaneous interaction, where ρ in [0,1] is a restitution factor. and represent the velocity of the system before and after the instant of impacts; S0 denotes the position of the nonzero rigid barrier.

The constitutive model of viscoelastic damping is provided as[11] and the relaxation function q(t) is expressed as follows based on Refs. [9] and [11]: Each λi in Eq. (3) is the relaxation time of the viscoelastic components, and βi specifies elastic modulus being either positive or negative. Both parameters are determined by the concrete problems.

2.2. Treatment for the viscoelastic force

Taking into account that the viscoelastic force contributes to both the stiffness and the damping,[16] system equations (1a) and (1b) can be written as Here, the consideration of the viscoelastic force has been incorporated in and r1 (x), respectively. The potential energy function U(x) and the total energy E of the system are The period of the corresponding free motion is given by where A is the amplitude for a given E, and the positive root is calculated from E = U(A).

The motion equations of the system (4a) and (4b) can be described through variable substitution. Utilize the total energy E(t) and the total phase angle ϑ(t) according to the transformation and ϑ(t) is defined as ϕ (t) denotes the residue phase angle, and ω(t) is the instantaneous frequency. The average frequency of ω(t) can be calculated by the quasi-period T Denote each component in the viscoelastic force exhibited in Eqs. (2) and (3) as Applying the transformation of Eq. (8) with , and neglecting the decaying transient part, we can get Thus, functions and r1 (x) in Eq. (4a) are derived by Eq. (12) as Besides, combining Eqs. (7) and (10), the average frequency is determined Thus, viscoelastic system equations (1a) and (1b) with unilateral nonzero offset barrier impacts are transformed to the system Eqs. (4a) and (4b) without viscoelastic damping form.

3. Stochastic averaging approach
3.1. Free VI system analysis

Introducing , equation (4a) can be rewritten as the following Itô stochastic differential equations: where Γ(x,y) and Σ(x,y) denote drift and diffusion representations, and W(t) is a unit Weiner process.

The potential energy function and total energy of the system have been reported in the previous Eqs. (5) and (6), respectively. When the parameter ε is zero, equations (15) and (4b) constitute a free VI system. Suppose functions r1 (x) and U(x) are such that the free VI system associated with Eq. (15) without impact has a gens periodic solutions surrounding the origin on phase plane (x,y). The period is the same as earlier Here, a denotes the maximum displacement derived by E = U(a). As E > U(S0), the system will impact the nonzero offset barrier. The system state before impact is , and the system energy can be written as When the impact occurs, the system state after impact with the nonzero offset barrier turns into , and the system energy becomes The loss of the system energy during each impact is calculated as The system still moves as a free oscillator between the first impact (t = t1) and the second impact (t = t2), and the time period Δt of the process between both consecutive impacts can be evaluated by The time period can be seen as the period of free oscillation after occurrence of impacts, which relies on the system energy; thus, denoted as T(E) = Δt represents the quasi-period. It is implied from Eq. (19) that the energy loss ΔE is related to the restitution factor ρ, the energy before nonzero rigid barrier impacts, the nonzero position S0, and the viscoelastic force due to its equivalent stiffness part.

3.2. Stochastic averaging of energy envelope (SAEE)

Relying on the Eqs. (6) and (15), the Itô equations for displacement x and system energy E can be derived by employing the Itô differential rule. When the damping, the stochastic excitation, and the energy loss are small, it can be found that energy E(t) is a slowly varying process, and the displacement x(t) is a rapidly varying process. Thus, the method of SAEE[45] is absorbed here. Further, the averaged Itô equation for E(t) is given as where Γ(E) and Σ(E) are averaged drift and diffusion representations, and B(t) is the standard Brownian motion process.

For the situation of E < U(S0), the system keeps the vibration without impact. The drift and diffusion representations of the viscoelastic system with nonzero offset barrier impacts remain as As E exceeds U(S0), the system possesses the vibration with the impact. The energy loss during impact needs to be counted; thus, the averaged drift and diffusion representation becomes where A = U−1(E) denotes the amplitude of vibration, and ΔE[T(E)]−1 describes the energy loss during each impact.

Further, the averaged Fokker–Plank–Kolmogorov (FPK) equation associated with Itô Eq. (21) is expressed as where p = p(E, t|E0) represents the transition probability density of E dependent on the initial condition and the boundary condition is finite for the situation of E = 0, while Thus, the stationary solution of Eq. (24) can be provided by where Z is the normalization constant.

Next, the joint probability density of the displacement and the velocity can be calculated Here, p(x|E) is the conditional probability density function, which can be given by The normalization constant Z′ is computed by Substituting Eqs. (29) and (30) into Eq. (28) yields Then, the marginal probability density functions p(x) and p(y) can be solved by

4. Application and simulation

Here, we adopt the nonlinear viscoelastic oscillator with right nonzero rigid barrier impacts to illustrate the analytical procedure The damping force and viscoelastic force V′ here are assumed to be of order ε, and η(t) are stationary broad-band random processes of order ε1/2, i.e., the Gaussian white noise with intensity 2D′. σ denotes the nature frequency. ρ and S0 are described the same as the previous interpretation. By approximating V′ based on Eqs. (12) and (13) to the case, the equivalent system of Eq. (33) has the form where and .

According to Eq. (14), specifically for the present case, the average frequency can be given by Relying on the method of SAEE, the drift and diffusion representations of the averaged Itô equation for energy E can be obtained in detail.

For the situation of E < U(S0), combining with Eqs. (22a)–(22c), let , and then yield

As E exceeds U(S0), combining with Eqs. (23a)–(23c), we can get

Next, by solving the averaged FPK Eq. (24), we obtain the PDFs of the system energy E, and further get the probability density functions p(x,y), p(x), and p(y) based on Eqs. (31), (32a), and (32b). And then consider the case of , system equation (33) represents a nonlinearly damped and stochastically excited Van der Pol viscoelastic VI system.

4.1. Validity of the approach

The analytical stationary PDFs will be exhibited and compared with the directly numerical simulation results. All the directly numerical simulations computed by employing the Runge–Kutta (RK) algorithm and Monte Carlo method are presented based on the original system Eq. (33), which provide the validation as well as rationality of the analytic results. Solutions are initially observed with the parameters: d0 = −0.01, d2 = −0.005, σ = 1.0, β1 = −0.2, λ1 = 1.1, ρ = 0.9, S0 = 0.995, and D′ = 0.0236, unless otherwise stated. Then, the effects on the system responses for the intensity D′ of stochastic process η(t), two viscoelastic parameters β1 and λ1, the position S0 of the right nonzero rigid barrier, and the restitution factor are described in detail in Figs. 25. The increase of noise intensity D′, i.e., gaining stronger external force leads to lower peaks of stationary PDFs of the displacement x and the velocity y, and smaller values for the stationary PDFs of total energy E at small energy, i.e., make ones more flat in Figs. 2(a)2(c) and the oscillator stays at equilibrium with a smaller possibility. Besides, figure 2(a) shows that the increase of D′ can contribute to the system with high energy, large displacement, and velocity, i.e., obtaining larger probability of the system. It can be seen that the approximate analytical results are relatively close to the directly numerical ones.

Fig. 2. Stationary PDFs of (a) total energy E, (b) displacement x, and (c) velocity y, solved both analytically and numerically with different noise intensities. Solid lines represent analytical results, while circles, squares, and triangles represent Monte Carlo simulation results.
Fig. 3. Stationary PDFs of (a) total energy E, (b) displacement x, and (c) velocity y, solved both analytically and numerically with different viscoelastic parameters. Solid lines represent analytical results, while circles, squares, and triangles represent Monte Carlo simulation results.
Fig. 4. Stationary PDFs of (a) total energy E, (b) displacement x, and (c) velocity y, solved both analytically and numerically with different positions of the nonzero rigid barrier. Solid lines represent analytical results, while circles, squares, and triangles represent Monte Carlo simulation results.
Fig. 5. Stationary PDFs of total energy (a) E, (b) displacement x, and (c) velocity y, solved both analytically and numerically with different restitution factor values. Solid lines represent analytical results, while circles and squares represent Monte Carlo simulation results.

The elastic modulus β1 and relaxation time λ1 show a difference in responses of the system according to Fig. 3. β1 keeps negative in Fig. 3, and the change of stationary PDFs caused by the variation of |β1| is stronger than that caused by the variation of λ1. The increase in |β1| is equivalent to an increase in the damping effect, which can contribute to higher peaks of stationary PDFs of displacement and velocity, and larger values for the stationary PDFs of total energy E at small energy, indicating a larger probability that the system will stay close to the equilibrium state. The two solutions here also agree well.

Examination of the position S0 of the right nonzero rigid barrier and the restitution factor ρ on the system response are illustrated in Figs. 4 and 5, respectively. Curves in Fig. 4(a) show unsmoothness at E = U(S0). When the oscillator is further away from the obstacle position in Figs. 4(b) and 4(c), which is equivalent to reducing the damping effect, it leads to lower peaks of stationary PDFs of displacement and velocity, and smaller values for the stationary PDFs of total energy E at small energy, indicating a smaller probability that the system will stay close to the equilibrium state. Additionally, when S0 reaches the amplitude of the oscillator motion, i.e., the maximum displacement which is approximately 1.995 here, the corresponding response probability is close to 0, indicating that if S0 crosses the threshold, the impacts will not occur. In Figs. 5(a)5(c), the variation of restitution factor ρ from 0.45 to 0.94 does not cause a great change in the response probability of the system. It can be found that the analytical results and MC simulation results agree well in these figures, which further indicates that the analytical procedure is reliable and effective via the related comparisons.

4.2. Further analysis: stochastic P-bifurcations

According to Section 4.1, satisfactory agreement is shown between the results of analytical procedure and MC simulation solution obtained from the original stochastic VI system. Based upon the above conclusion, a further analysis for joint stationary PDFs of the displacement x and the velocity y is observed here, and stochastic P-bifurcation occurs with the variation of certain parameters. Thus, we perform the exploration for stochastic P-bifurcation in this subsection. Take ρ = 0.95 and S0 = 0.985, and the rest are kept the same as the initial values.

The position S0 of the nonzero rigid barrier varies from 0.985 in Figs. 6(a), 6(b), and 6(d) to 1.985 in Fig. 6(c), and the figures exhibit the decreasing joint PDFs and the corresponding analytical and numerical contour lines for different elastic moduli β1 being negative, which is the response to the result in Fig. 3. The contours of the two solutions further change from Fig. 7(a) to Fig. 7(d) with the large-amplitude oscillations and the small-amplitude oscillations around the origin at the position of S0 = 0.485 when some parameters are in the bifurcation schemes, and the phenomenon here will be discussed below. Besides, the analytical solutions and MC simulation ones on the contours of joint PDFs agree well here, which responds to the previous conclusion.

Fig. 6. The contours for joint PDFs of the displacement x and the velocity y solved both analytically (solid line) and numerically (pecked line) with different elastic modulus values β1: (a) β1 = −0.5, (b) β1 = −0.19, (c) β1 = −0.19, S0 = 1.985, and (d) β1 = −0.06.
Fig. 7. The contours for joint PDFs of the displacement x and the velocity y solved both analytically (solid line) and numerically (dotted line) with different bifurcation parameters: (a) β1 = 0.0, (b) λ1 = 0.01, (c) ρ = 1.0, and (d) d0 = −0.05.

Next, figures 8, 10, 12, and 14(a)14(c) provide joint stationary PDFs of the displacement x and the velocity y and the section graphs of the joint PDFs on the surface x = 0.485 for different damping coefficients d0, elastic moduli β1, relaxation times λ1, and restitution coefficients ρ, respectively. When ρ = 0.95, β1 = −0.1, and λ1 = 0.5 are selected, the joint PDFs decreases as |d0| increases, reflecting the effect of the reduced damping on the small displacement and velocity. In Fig. 8(c), it demonstrates the process of evolving from a unimodal shape of joint PDFs, which shows at the position of the rigid barrier, to a bimodal shape, i.e., the shape of crater in Fig. 8(b). Figures 9(a)9(d) further provide the corresponding contours. When d0 = −0.015 in Fig. 9(a), there exists one stochastic attractor, characterized by small-amplitude oscillations around the origin. And when d0 reaches −0.035, the stochastic attractor is transformed into the characterization of large-amplitude oscillations. When d0 is further changed to −0.045 and −0.05 in Figs. 9(c) and 9(d), the small-amplitude oscillations around the origin appear again and get stronger, while the existing large-amplitude oscillations is weakened. Clearly, it indicates that the occurrence of stochastic P-bifurcation exists in the process of variations of d0 based upon the concept of stochastic bifurcation (phenomenological or P-bifurcations are observed when the underlying probabilistic structure of the long-term behavior of the state variables undergoes topological changes with certain parameter variations),[4649] and it occurs when d0 is close to −0.029, and the stability of system is changed here.

Fig. 8. The surface for joint PDFs of the displacement x and the velocity y with the damping coefficient (a) d0 = −0.015 and (b) d0 = −0.05; (c) the sectional views of joint PDFs on the surface x = 0.485 with variations of the damping coefficient d0.
Fig. 9. Contour plots for joint PDFs of the displacement x and the velocity y: (a) d0 = −0.015, (b) d0 = −0.035, (c) d0 = −0.045, and (d) d0 = −0.05.
Fig. 10. The surface for joint PDFs of the displacement x and the velocity y with the elastic modulus (a) β1 = 0.0 and (b) β1 = 0.06; (c) the sectional views of joint PDFs on the surface x = 0.485 with variations of the elastic modulus β1.

When d0 returns to the initial value, a non-negative value of the elastic modulus β1 of viscoelastic force induces a similar phenomenon. Figures 10(a) and 10(b) illustrate the joint stationary PDFs with the shape of crater for β1 = 0.0 and β1 = 0.06, which are different from the effect in Fig. 6. It intuitively shows the evolution process of the joint PDFs from the unimodal shape to the shape of crater caused by varying β1 from negative to zero and then to positive (equivalent to reducing damping) through the section graphs of joint PDFs when x reaches the position of rigid barrier in Fig. 10(c). Hence, it can be seen that the stochastic P-bifurcation occurs when β1 approximates to −0.02, which is treated as the watershed of joint PDFs from the singular peak to the shape of crater. Figures 11(a)11(d) also report the corresponding contours. The variation of the stochastic attractor caused by varying β1 from negative to zero and then to positive is close to the effect of the increase of |d0| in Fig. 9.

Fig. 11. Contour plots for joint PDFs of the displacement x and the velocity y: (a) β1 = −0.21, (b) β1 = 0.0, (c) β1 = 0.03, and (d) β1 = 0.06.
Fig. 12. The surface for joint PDFs of the displacement x and the velocity y with the relaxation time (a) λ1 = 0.01 and (b) λ1 = 2.7; (c) the sectional views of joint PDFs on the surface x = 0.485 with variations of the relaxation time λ1.

Further, we examine the relaxation time λ1 of viscoelastic force from the vicinity of the usual value 1.0 to a wider range, and obtain interesting observations. Figures 12(a) and 12(b) exhibit joint stationary PDFs possessing the shape of crater for λ1 = 0.01 and λ1 = 2.7, which are different from the effect in Fig. 3. Figure 12(c) intuitively shows the evolution process of the joint PDFs from the unimodal shape to the shape of crater as λ1 varies from 0.01 to 0.37 and from 1.59 to 2.7 (equivalent to reducing damping) through the section views of joint PDFs when x reaches the position of rigid barrier. Clearly, it can be seen the stochastic P-bifurcation occurs when λ1 approximates to 0.13 or 1.87. The corresponding contours depicted in Figs. 13(a)13(d) show that the transformation on the characterization of the stochastic attractor from small-amplitude oscillations around the origin to large-amplitude oscillations indeed exists in the above two intervals, which has not been covered before.

Fig. 13. Contour plots for joint PDFs of the displacement x and the velocity y: (a) λ1 = 0.01, (b) λ1 = 0.5, (c) λ1 = 1.5, and (d) λ1 = 2.7.
Fig. 14. The surface for joint PDFs of the displacement x and the velocity y with the restitution coefficients (a) ρ = 0.917 and (b) ρ = 1.0; (c) the sectional views of joint PDFs on the surface x = 0.485 with variations of the restitution coefficients ρ.

Then choose β1 = −0.1 and λ1 = 0.65, we further consider the effect of the restitution factor ρ limited to the range 0.9–1.0 on the system response. The effect of impact is to increase the damping associated with the system, and qualitatively, decreasing ρ is equivalent to increasing the damping parameter. The stationary joint PDFs are computed for the cases ρ = 0.917 and 1.0, as shown in Figs. 14(a) and 14(b). It decreases for the small displacement and velocity when ρ reaches 1.0, representing the classical elastic impact, and the shape of the crater also appears in the figure. The section views of joint PDFs in Fig. 14(c) further describe the evolution process caused by ρ on the surface of rigid wall and report the watershed quickly, i.e., when ρ approximates to 0.968, the stochastic P-bifurcation occurs, and it becomes double peaks when ρ = 1.0, corresponding to the shape of the crater in Fig. 14(b). Figures 15(a)15(d) also provide the corresponding contour plots. The characterization of the stochastic attractors develops into large amplitude oscillations at ρ = 0.986 in Fig. 15(b), and both small amplitude oscillations and large amplitude oscillations are visible in Figs. 15(c) and 15(d).

Fig. 15. Contour plots for joint PDFs of the displacement x and the velocity y: (a) ρ = 0.917, (b) ρ = 0.986, (c) ρ = 0.9977, and (d) ρ = 1.0.
5. Conclusions

In this paper, stationary responses have been exhibited to characterize the vibration law of the stochastic viscoelastic system with right nonzero offset barrier impacts. First, the viscoelastic force is replaced by a combination of equivalent effect terms. The motions of free VI systems are sorted into periodic motion without impact and quasi-periodic motion with impact based on the levels of system energy. Then, the impact condition for the displacement and velocity is transformed into the system energy description. The stationary PDFs of system are then obtained by utilizing SAEE based upon the assumption of lightly damping and weakly random perturbation, and the accuracy of the method is substantiated by comparing the theoretical results with those from MC simulations via the variations of the noise intensity, the viscoelastic parameters, the position of right nonzero rigid wall, and the restitution factor. Through the presented approach, the stochastic P-bifurcation is explored, which occurs when any of the damping coefficient, the elastic modulus, the relaxation time, or the restitution factor arrives at the watershed of joint stationary PDFs from the unimodal shape to the shape of crater, and the system stability is changed here. It is interesting that the change of elastic modulus from negative to zero and then to positive, instead of just the negative value in the previous considerations, has witnessed the evolution process of stochastic P-bifurcation. From the vicinity of the usual value to a wider range, the relaxation time induces the stochastic P-bifurcation in the two interval schemes. The suggested procedure shows some advantages, including conciseness and operability. It is a significant study for seeking the viscoelastic property and the impact condition in the future.

Reference
[1] Mrlík M Ilcikova M Sedlacik M Mosnacek J Peer P Filip P 2014 Colloid Polym. Sci. 292 2137
[2] Arenas J P Crocker M J 2010 Sound Vib. 44 12
[3] Lu L M Wen J H Zhao H G Wen X S 2014 Acta Phys. Sin. 63 154301 in Chinese
[4] Kiik J C Kurasov P Usman M 2015 Phys. Lett. A 379 1871
[5] Khan Z El Naggar M Cascante G 2011 J. Franklin I. 348 1363
[6] Findley W N Davis F A 2013 Creep and Relaxation of Nonlinear Viscoelastic Materials Amsterdam North-Holland
[7] Del Nobile M Chillo S Mentana A Baiano A 2007 J. Food Eng. 78 978
[8] Renaud F Dion J L Chevallier G Tawfiq I Lemaire R 2011 Mech. Syst. Signal. Pr. 25 991
[9] Christensen R 2012 Theory of Viscoelasticity: an Introduction New York Academic Press
[10] Roscoe R 1950 Br. J. Appl. Phys. 1 171
[11] Drozdov A D 1998 Viscoelastic Structures: Mechanics of Growth and Aging San Diego Academic Press
[12] de Espíndola J J Bavastri C A Lopes E M 2010 J. Franklin I. 347 102
[13] Han X Wang M 2009 Math. Methods Appl. Sci. 32 346
[14] Zhu W Q 1992 Random Vibration Beijing Science Press in Chinese
[15] Potapov V 1997 J. Appl. Math. Mech. 61 287
[16] Zhu W Q Cai G Q 2011 Internat. J. Non-Linear Mech. 46 720
[17] Huang Q H Xie W C 2008 J. Appl. Mech. 75 021012
[18] Di Paola M Failla G Pirrotta A 2012 Probabilist. Eng. Mech. 28 85
[19] Soize C Poloskov I E 2012 Comput. Math. Appl. 64 3594
[20] Zhu WQ 1996 Appl. Mech. Rev. 49 S72
[21] Brogliato B 1996 Nonsmooth Impact Mechanics: Models, Dynamics and Control, LNCIS 220 Berlin Springer Verlag
[22] Shaw S Holmes P 1983 J. Appl. Mech. 50 849
[23] Dankowicz H Nordmark A B 2000 Phys. D 136 280
[24] Dimentberg M Iourtchenko D 2004 Nonlinear Dynam. 36 229
[25] Feng Q 2003 Comput. Methods Appl. Mech. Eng. 192 2339
[26] Feng Q He H 2003 Eur. J. Mech. A Solids 22 267
[27] Namachchivaya N S Park J H 2005 J. Appl. Mech. 72 862
[28] Park J H Namachchivaya N S 2004 ASME 2004 International Mechanical Engineering Congress and Exposition November 13-19, 2004 Anaheim, USA 189 200
[29] Zhuravlev V F 1976 Mech. Solids 11 23
[30] Wu Y Zhu W Q 2008 Phys. Lett. A 372 623
[31] Feng J Xu W Wang R 2008 J. Sound Vib. 309 730
[32] Dimentberg M Gaidai O Naess A 2009 Internat. J. Non-Linear Mech. 44 791
[33] Wang D L Xu W Zhao X R 2016 Chaos 26 033103
[34] Feng J Q Xu W Niu Y J 2010 Acta Phys. Sin. 59 157 in Chinese
[35] Kumar P Narayanan S Gupta S 2014 Probabilist. Eng. Mech. 38 143
[36] Kumar P Narayanan S Gupta S 2016 Nonlinear Dynam. 85 439
[37] Kumar P Narayanan S Gupta S 2017 Int. J. Mech. Sci. 127 103
[38] Su M B Rong H W 2011 Chin. Phys. B 20 060501
[39] Rong H W Wang X D Xu W Fang T 2005 Acta Phys. Sin. 54 4610 in Chinese
[40] Ma S J Xu W Li W 2006 Acta Phys. Sin. 55 4013 in Chinese
[41] Yang L H Ge Y Ma X K 2017 Acta Phys. Sin. 66 190501 in Chinese
[42] Li W Zhang M T Zhao J F 2017 Chin. Phys. B 26 090501
[43] Kumar P Narayanan S Gupta S 2016 Probabilist. Eng. Mech. 45 70
[44] Kumar P Narayanan S Gupta S 2016 Procedia Eng. 144 998
[45] Zhu W Q Lin Y K 1991 J. Eng. Mech. 117 1890
[46] Xu W He Q Rong H W Fang T 2003 Proceedings of the Fifth International Conference on Stochastic Structural Dynamics–SSD03 Zhu W Q Boca Raton CRC Press 509 515
[47] Dtchetgnia Djeundam S Yamapi R Kofane T Aziz-Alaoui M 2013 Chaos 23 033125
[48] Arnold L 2013 Random Dynamical Systems Berlin Springer-Verlag
[49] Zakharova A Feoktistov A Vadivasova T Schöll E 2013 Eur. Phys. J. Spec. TOP. 222 2481