• Aucun résultat trouvé

Hysteresis suppression for primary and subharmonic 3:1 resonances using fast excitation

N/A
N/A
Protected

Academic year: 2021

Partager "Hysteresis suppression for primary and subharmonic 3:1 resonances using fast excitation"

Copied!
13
0
0

Texte intégral

(1)

DOI 10.1007/s11071-008-9438-z

O R I G I N A L PA P E R

Hysteresis suppression for primary and subharmonic 3:1 resonances using fast excitation

Mohamed Belhaq·Abdelhak Fahsi

Received: 9 June 2008 / Accepted: 17 September 2008 / Published online: 9 October 2008

© Springer Science+Business Media B.V. 2008

Abstract We analyze the effect of a fast harmonic

excitation on hysteresis and on entrainment area in a forced van der Pol–Duffing oscillator near the primary and the 3:1 subharmonic resonances. Analytical treat- ment based on perturbation techniques is performed to capture the entrainment zone, the quasiperiodic mod- ulation domain and the hysteresis area in the vicin- ity of the two resonances. Specifically, it is shown that a fast harmonic excitation can suppress hystere- sis for a certain range of the fast excitation leading to a smooth transition between the quasiperiodic and the frequency-locked responses near these resonances.

Furthermore, the influence of different system para- meters on the hysteresis area has been investigated. In particular, the results reveal that the amplitude of the fast excitation and the nonlinear damping significantly affect the domain of hysteresis suppression near the primary and the 3:1 subharmonic resonances.

Keywords Fast harmonic excitation

· Frequency locking · Hysteresis suppression · Subharmonic resonance · Quasiperiodic modulation · Perturbation analysis

M. Belhaq (

)

University Hassan II-Aïn Chock, Casablanca, Morocco e-mail:mbelhaq@hotmail.com

A. Fahsi

FSTM, University Hassan II-Mohammadia, Mohammadia, Morocco

1 Introduction

Frequency locking behavior can occur in various me- chanical systems. In a such phenomenon, self-excited oscillations synchronize by periodic forcing of the sys- tem leading to frequency-locked oscillations for which the response follows the forcing frequency. Out of the resonance but nearby, the response is quasiperiodic.

In nonlinear systems, the transition between quasiperi- odic and entrained motions can take place at two dif- ferent specific frequencies when the forcing frequency is swept backward and forward leading to hysteresis effect. This effect, producing jumps in the system re- sponse is considered as one of the serious problems in the design of devices like resonant microsensors [1,

2].

Hysteresis occurring inside entrainment area, referred to as entrained hysteresis, has been reported and an- alyzed in optically driven MEMS resonators; see [3]

and references therein. The presence of such a hystere-

sis near a resonance is associated with the coexistence

of two distinct stable states (periodic and quasiperi-

odic) and, hence, a jump between these states may oc-

cur. Therefore, the control of hysteresis is an important

issue to realize a high functionality of systems and to

improve their specific performance. In [4], the control

of hysteresis in a directly modulated semiconductor

laser using delayed optoelectronic feedback was pro-

posed. In [2], the attenuation of hysteresis in MEMS

resonators was performed by acting on the quality fac-

tor which is related to the damping.

(2)

In a recent work [5], the suppression of hysteresis in a forced van der Pol–Duffing oscillator was studied near the fundamental resonance 1:1. It was shown that adding a fast harmonic excitation (FHE) can suppress hysteresis for a certain range of the fast frequency.

The purpose of the present paper is twofold. First, further investigation will be carried out for the forced van der Pol–Duffing oscillator near the 1:1 resonance.

In particular, the influence of different system parame- ters on the suppression of hysteresis will be examined.

Second, we perform a careful analysis of the effect of an FHE on frequency locking and on hysteresis sup- pression near the subharmonic resonance 3:1.

The idea of using FHE to study the stabilization of an inverted pendulum on a vibrating support has been reported in [6–8]. Other effects of FHE on mechani- cal systems have been examined intensively in the re- cent years. These include equilibrium stability [9], lin- ear stiffness [10], natural frequencies [11], resonance behavior [12], symmetry breaking [13], and limit cy- cle [14–16].

Consider the following forced van der Pol–Duffing oscillator subjected to an FHE in the dimensionless form:

¨ x + x

αβx

2

˙ xγ x

3

= h cos ωt +

2

cos x cos Ωt (1) where damping α, β , nonlinearity γ and excitation amplitudes h and a are small. Dots denote differen- tiation with respect to time t . We assume that the FHE frequency Ω is high compared to the frequency of the external forcing ω such that the resonance between the two frequencies cannot occur. In a previous work [17], a van der Pol–Mathieu–Duffing equation was inves- tigated near the 2:1 and the 1:1 resonances. It was shown that varying the FHE shifts the backbone curve and changes the nonlinear characteristic behavior of the system near these resonances from softening to hardening or vice versa.

In the present work, we focus our attention on the effect of an FHE on the frequency locking area and on the suppression of the entrained hysteresis in (1) near the fundamental 1:1 and the subharmonic 3:1 res- onances.

The rest of the paper is organized as follows. In Sect.

2

we average the oscillator (1) over the fast time to derive an equation governing the slow dy- namics. Section

3

is devoted to the analysis near

the 1:1 resonance. The multiple scales method is ap- plied to the slow dynamics to derive an autonomous slow flow. Analysis of equilibria of this slow flow provides analytical approximations of the entrained amplitude-frequency response. A multiple scales meth- od is performed in a second step on the slow flow to approximate quasiperiodic solution and its modula- tion domain. Analytical prediction of the variation of the hysteresis area as a function of the high frequency Ω as well as the influence of different parameters of the system on the hysteresis suppression is provided.

We perform numerical simulation and we compare with the analytical finding for validation. In Sect.

4,

we perform a similar analysis near the 3:1 resonance.

Section

5

concludes the work.

2 Slow dynamics

In this section, we use the method of direct partition of motion (DPM) [18] to derive the slow dynamics of system (1). We introduce two different time scales, a fast time T

0

= Ωt and a slow time T

1

= t , and we split up x(t ) into a slow part z(T

1

) and a fast part φ (T

0

, T

1

) as follows:

x(t ) = z(T

1

) + φ (T

0

, T

1

) (2) where z describes slow main motions at time scale of oscillations, φ stands for an overlay of the fast mo- tions and indicates that φ is small compared to z.

Since Ω is considered as a large parameter, we choose Ω

1

, for convenience. The fast part φ and its derivatives are assumed to be 2π-periodic functions of fast time T

0

with zero mean value with respect to this time, so that x(t ) = z(T

1

) where ≡

1

0

() dT

0

defines time-averaging operator over one period of the fast excitation with the slow time T

1

fixed.

Introducing D

ij

j

∂Tij

yields

dtd

= ΩD

0

+ D

1

,

d2

dt2

= Ω

2

D

02

+ 2ΩD

0

D

1

+ D

21

, and substituting (2) into (1) gives

1

D

20

φ + 2D

0

D

1

φ + D

21

φ + D

21

z + z + φ

αβ(z + φ)

2

(D

0

φ + D

1

φ + D

1

z)

γ (z + φ)

3

= h cos ωT

1

+

1

(aΩ) cos(z + φ) cos T

0

. (3) Averaging (3) leads to

D

21

z + z

αβz

2

D

1

zγ z

3

= h cos ωT

1

+

1

(aΩ)

cos(z + φ) cos T

0

. (4)

(3)

Subtracting (4) from (3) yields

1

D

02

φ + 2D

0

D

1

φ + D

21

φ + φ

αβz

2

(D

0

φ + D

1

φ) + β

2φ +

2

φ

2

(D

0

φ + D

1

φ)

γ

3z

2

φ + 3

2

2

+

3

φ

3

=

1

(aΩ) cos(z + φ) cos T

0

1

(aΩ)

cos(z + φ) cos T

0

. (5)

Approximation of φ is obtained from (5) by consid- ering only the dominant terms of order

1

as D

20

φ = (aΩ) cos z cos T

0

(6) where it is assumed that = O(

0

). The stationary solution to the first order for φ is written as

φ = −a cos z cos T

0

. (7)

The equation governing the slow motion is derived from (4). Inserting cos(z + φ) = cos zφ sin z + O(

2

) into (4) and retaining the dominant terms of or- der

0

, we obtain

D

21

z + z

αβz

2

D

1

zγ z

3

= h cos ωT

1

(aΩ)sin z φ cos T

0

. (8) Inserting φ from (7) and using that cos

2

T

0

= 1/2, we find the approximate equation for slow motions D

21

z + z

αβz

2

D

1

zγ z

3

= h cos ωT

1

+ 1

2 (aΩ)

2

cos z sin z. (9) This equation is similar to the original equation (1) in which the non-autonomous term

2

cos x cos Ωt is replaced by the autonomous one

12

(aΩ)

2

cos z sin z.

We focus the analysis on small vibrations around the origin by expanding in Taylor’s series the terms sin z zz

3

/6 and cosz 1 − z

2

/2. Keeping only terms up to order three in z, (9) becomes

D

21

z +

1 − 1 2 (aΩ)

2

z

αβz

2

D

1

z

γ − 1 3 (aΩ)

2

z

3

= h cos ωT

1

. (10)

In (10), it appears that the effect of an FHE introduces additional apparent stiffness in the linear stiffness [9]

and in the nonlinear one. These effects have been ob- served in a spring-connected two-link mechanism at a vibrating support [12].

Hence, (10) can be written as

¨

z + ω

20

z

αβz

2

˙

zξ z

3

= h cos ωt (11) where ω

20

= 1 −

12

(aΩ)

2

, ξ = γ

13

(aΩ)

2

and an overdot denotes differentiation with respect to time t .

3 The fundamental resonance 1:1

We express the 1:1 resonance condition by introducing a detuning parameter σ according to

ω

20

= ω

2

+ σ. (12)

Analytical investigation of periodic and quasiperiodic responses near the resonance requires the application of a double perturbation technique by introducing two small bookkeeping parameters, μ and η. To derive a slow flow of the slow dynamics (11), we use the para- meter μ in the first perturbation, and for implementing the second perturbation on the slow flow, we introduce the other parameter η in a second step. This strategy of using a perturbation analysis on a slow flow in systems under quasiperiodic excitation has been proposed in [19] and applied successfully to obtain analytical ap- proximations of quasiperiodic solutions [20] and ana- lytical expressions of the stability chart for quasiperi- odic Mathieu equations [21–23].

3.1 Slow flow and entrainment Rewrite (11) as

¨ z + ω

2

z

= μσ z +

αβz

2

˙

z + ξ z

3

+ h cos ωt

. (13) Using the multiple scales technique [24], we seek a two-scale expansion of the solution in the form z(t ) = z

0

(T

0

, T

1

) + μz

1

(T

0

, T

1

) + O

μ

2

(14) where T

i

= μ

i

t . In terms of the variables T

i

, the time derivatives become

dtd

= D

0

+ μD

1

+ O(μ

2

) and

d2

dt2

= D

02

+2μD

0

D

1

+ O(μ

2

), where D

ji

=

j

∂Tij

. Sub-

(4)

Fig. 1 Amplitude-frequency response for different values ofΩ. Analytical approximation: solid (for stable) and dashed (for unstable).

Numerical simulation: circles

stituting (14) into (13) and equating coefficients of the same power of μ, we obtain

D

20

z

0

(T

0

, T

1

) + ω

2

z

0

(T

0

, T

1

) = 0, (15) D

20

z

1

(T

0

, T

1

) + ω

2

z

1

(T

0

, T

1

)

= −2D

0

D

1

z

0

σ z

0

+

αβz

20

D

0

z

0

+ ξ z

03

+ h cos ωT

0

. (16) The solution to the first order is given by

z

0

(T

0

, T

1

) = r(T

1

) cos

ωT

0

+ θ (T

1

)

. (17)

Substituting (17) into (16), removing secular terms and using the expressions

drdt

= μD

1

r + O(μ

2

) and

dt

= μD

1

θ + O(μ

2

), we obtain the first-order au- tonomous slow flow modulation equations of ampli- tude and phase

dr

dt = ArBr

3

H sin θ, r

dt = SrCr

3

H cos θ,

(18)

where A =

α2

, B =

β8

, S =

σ

, C =

and H =

h

. Equilibrium points of the slow flow (18), correspond- ing to periodic solutions of (11), are determined by setting

drdt

=

dt

= 0. Using the relation cos

2

θ + sin

2

θ = 1, we obtain the amplitude-frequency re- sponse equation

A

3

r

6

+ A

2

r

4

+ A

1

r

2

+ A

0

= 0 (19)

where A

3

= B

2

+ C

2

, A

2

= −2(AB + SC), A

1

= A

2

+ S

2

and A

0

= − H

2

. The discriminant of (19) is given by

= P

3

27 + Q

2

4 (20)

where P =

AA13

3AA222 3

and Q =

272

(

AA2

3

)

3

A3A2A21 3

+

AA03

. Equation (19) has three real positive roots if is neg- ative. Furthermore, (19) has only one positive root if is positive. In what follows, we fix the parameters α = 0.01, β = 0.05, γ = 0.1, h = 0.1 and a = 0.02. In Fig.

1(a), we show the frequency-response curve, as

given by (19), for Ω = 0. The solid lines denote stable branches and the dashed lines denote unstable ones.

The effect of the frequency Ω on the backbone curve is illustrated in Fig.

1(b), (c) for the values

Ω = 25 and Ω = 40, respectively. The plots show that as Ω increases, the backbone curve shifts left and its char- acteristic switches from softening to hardening. For validation, analytical approximations are compared to numerical integration (circles) using a Runge–Kutta method.

3.2 Quasiperiodic modulation and hysteresis suppression

We transform the polar form (18) using the variable

change u = r cos θ, v = − r sin θ , and we implement a

second perturbation analysis by introducing the book-

keeping parameter η in damping and in nonlinearity

(5)

components. We obtain the Cartesian system du

dt = Sv + η Au(Bu + Cv)

u

2

+ v

2

, dv

dt = − Su + H

+ η Av(BvCu)

u

2

+ v

2

.

(21)

Approximations of periodic solutions of the slow flow (21), corresponding to quasiperiodic motion of the slow dynamics (11), can be obtained by using a multi- ple scales technique [19,

21]. We expand solutions as

u(t ) = u

0

(T

0

, T

1

) + ηu

1

(T

0

, T

1

) + O η

2

, v(t ) = v

0

(T

0

, T

1

) + ηv

1

(T

0

, T

1

) + O

η

2

,

(22)

where T

i

= η

i

t. Introducing D

i

=

∂Ti

yields

dtd

= D

0

+ ηD

1

+ O(η

2

), substituting (22) into (21) and col- lecting terms, we get:

– Order η

0

:

D

20

u

0

+ S

2

u

0

= SH,

Sv

0

= D

0

u

0

; (23)

– Order η

1

: D

02

u

1

+ S

2

u

1

= S

D

1

v

0

+ Av

0

(Bv

0

Cu

0

)

u

20

+ v

20

D

0

D

1

u

0

+ AD

0

u

0

D

0

Bu

0

+ Cv

0

u

20

+ v

02

, Sv

1

= D

0

u

1

+ D

1

u

0

Au

0

+ (Bu

0

+ Cv

0

)

u

20

+ v

02

.

(24)

The solution to the first-order system (23) is given by u

0

(T

0

, T

1

) = H

S + R(T

1

) cos

ST

0

+ ϕ(T

1

) , v

0

(T

0

, T

1

) = − R(T

1

) sin

ST

0

+ ϕ(T

1

) .

(25)

Substituting (25) into (24) and removing secular terms gives the following autonomous slow slow-flow sys- tem on R and ϕ:

dR dt =

A − 2BH

2

S

2

RBR

3

,

dt = − CR

2

− 2CH

2

S

2

.

(26)

Then, the approximate periodic solution of the slow flow (21) is given by

u(t ) = H

S + R cos νt, v(t ) = −R sin νt,

(27)

where the amplitude R is obtained by setting

dRdt

= 0 and given by

R =

A B − 2H

2

S

2

, (28)

and the frequency ν (frequency of the slow-flow limit cycle) is given by

ν = SCR

2

− 2CH

2

S

2

, (29)

and hence the approximate quasiperiodic response of the slow dynamics (11) reads

z(t ) = H

S cos ωt + R cos(ω + ν)t. (30) On the other hand, using u(t ) = r(t ) cos θ (t ), v(t ) =

r(t )sin θ (t ), the modulated amplitude r(t ) of the quasiperiodic oscillations is approximated by

r(t ) =

A BH

2

S

2

+ 2R H

S cos νt . (31) The envelope of this modulated amplitude is delimited by r

min

and r

max

given by

rmin=min

A

BH2 S2 +2RH

S,

A BH2

S2 −2RH S

,

(32)

rmax=max

A

BH2 S2 +2RH

S,

A BH2

S2 −2RH S

.

(33)

In Fig.

2

we draw, for Ω = 25, the modulated qua-

siperiodic area given by (32), (33) (solid lines) and

obtained numerically using Runge–Kutta method (cir-

cles). The comparison between the two results shows

a good agreement and clearly confirms the accuracy

of the analytical approach used here. In contrast, the

approach used in [17], based on the invariance of the

slow flow under the transformation θ → − θ +

π2

,

(6)

σ → − σ and ξ → − ξ , presented a slight discrepancy between the theory and the simulation, and fails in the vicinity of jumps phenomena, especially. Figure

2

shows that outside the synchronization area, quasiperi- odic behavior with two predominant frequencies takes place. Moving away from the synchronization area, the depth of modulation amplitude decreases. When approaching the entrainment area, the lower band of the modulation amplitude domain drops to zero, and then increases to collide with the upper limit of the modulation amplitude, exactly on the loci where the frequency-locked response takes place. Note that this dynamics has not been captured analytically using the invariance approach [17].

Fig. 2 Modulation amplitude vibration forΩ=25. Analytical approximation: solid lines. Numerical simulation: circles. QP:

modulation area of quasiperiodic response

In Fig.

3(a), (b) we show a global picture including

the quasiperiodic modulation area and the amplitude- frequency response for the values Ω = 0 and Ω = 40, respectively. Figure

3(c) illustrates the frequency re-

sponse for the critical case Ω

c

corresponding to the vanishing of the nonlinear stiffness component ξ of the slow dynamics (11). This threshold is given by Ω

c

=

√ 3γ

a . (34)

By analyzing the sign of the discriminant given by (20), we obtain an analytical approximation of the hys- teresis width as illustrated in Fig.

3(a), (b). In Fig.4,

we plot this hysteresis area as a function of the fast ex- citation frequency Ω . It can be seen from Fig.

4

that a complete elimination of the hysteresis is achieved in a certain range of the frequency Ω around the threshold Ω

c

27.4 located exactly in the middle of the sup- pression domain. In this domain, jump phenomenon is completely eliminated and hence a smooth transition between the quasiperiodic response and the frequency- locked motion takes place. This analytical prediction illustrated in Fig.

4

is confirmed by comparison to nu- merical simulations provided in Fig.

1.

In Fig.

5, we plot the effect of different system pa-

rameters on hysteresis. When varying a parameter, the others are kept fixed with the values as given above.

Figure

5(a) shows the effect of varying the amplitude

of the external forcing h on the hysteresis area and on the hysteresis suppression domain. It can be seen that for increasing values of h, the hysteresis area in- creases, whereas the suppression domain is still un- changed. This indicates that the external forcing am-

Fig. 3 Effect of the high frequencyΩon the hysteresis width. H: hysteresis area; E: entrainment area

(7)

plitude h has no effect on the hysteresis suppression area. Figure

5(b) shows the effect of varying the linear

damping α on the hysteresis. This subfigure indicates that for the given parameters, the linear damping has

Fig. 4 Variation of the area of the hysteresis loop versus the frequency excitationΩ. Analytical result from (20)

no effect on the hysteresis suppression domain and a small influence on the hysteresis area. Figure

5(c) in-

dicates that increasing nonlinear damping β , the hys- teresis area decreases and the suppression domain in- creases significantly. It can be seen from Fig.

5(c) that

for higher values of β , hysteresis can be suppressed from lower values of Ω . Finally, Fig.

5(d) illustrates

the effect of the amplitude a of the FHE on the hystere- sis. The plots reveal that as a is increased, the suppres- sion domain decreases and shifts left toward smaller values of Ω. This suggests that adjustment of the sup- pression area to a desired frequency range of Ω may be achieved by acting on the amplitude of the FHE.

4 The subharmonic resonance 3:1

In this section, we analyze the effect of an FHE on frequency locking and on hysteresis near the subhar- monic resonance 3:1. We express this resonance con- dition by introducing the detuning parameter σ ac-

Fig. 5 (Color online)

Effect of system parameters on hysteresis area and on hysteresis suppression domain. (a) Effect of the external excitationh, (b) effect of the linear dampingα, (c) effect of the nonlinear dampingβ, (d) effect of the amplitudea of the FHE

(8)

cording to ω

20

=

ω 3

2

+ σ. (35)

Analytical treatment of periodic and quasiperiodic so- lutions is carried out as in the previous case of reso- nance 1:1.

4.1 Slow flow and entrainment

The first perturbation step to derive a slow flow is im- plemented using the parameter μ. According to (35), rewrite (11) as

¨ z +

ω 3

2

z

= h cos ωt + μσ z +

αβz

2

˙ z + ξ z

3

. (36) We seek a two-scale expansion of the solution in the form

z(t ) = z

0

(T

0

, T

1

) + μz

1

(T

0

, T

1

) + O μ

2

(37) where T

i

= μ

i

t . Substituting (37) into (36) and equat- ing coefficients of the same power of μ, we obtain a set of linear partial differential equations

D

02

z

0

(T

0

, T

1

) + ω

3

2

z

0

(T

0

, T

1

) = h cos ωT

0

, (38) D

02

z

1

(T

0

, T

1

) +

ω 3

2

z

1

(T

0

, T

1

)

= −2D

0

D

1

z

0

σ z

0

+

αβz

20

D

0

z

0

+ ξ z

30

. (39) The solution to the first order is given by

z

0

(T

0

, T

1

)

= r(T

1

)cos ω

3 T

0

+ θ (T

1

)

+ F cos ωT

0

. (40) Substituting (40) into (39), removing secular terms and using the expressions

drdt

= μD

1

r + O(μ

2

) and

dt

= μD

1

θ + O(μ

2

), we obtain to the first order the autonomous slow-flow modulation equations of am- plitude and phase:

dr

dt = ArBr

3

(H

1

sin 3θ + H

2

cos 3θ )r

2

,

dt = SCr

2

(H

1

cos 3θ − H

2

sin 3θ )r,

(41)

where A =

α2

βF42

, B =

β8

, C =

, S =

9ξ F2

, H

1

=

9ξ F

, H

2

=

βF8

and F = −

9h2

.

Equilibria of the slow flow (41), corresponding to pe- riodic oscillations of (11), are determined by setting

dr

dt

=

dt

= 0. We obtain the amplitude-frequency re- sponse equation

A

2

r

4

+ A

1

r

2

+ A

0

= 0 (42) and the phase-frequency response relation

tan 3θ = (ABr

2

)H

1

(SCr

2

)H

2

(ABr

2

)H

2

+ (SCr

2

)H

1

(43) where A

2

= B

2

+ C

2

, A

1

= − (2AB + 2SC + H

12

+ H

22

) and A

0

= A

2

+ S

2

.

Figure

6(a), (b) shows the variation of

z(t )-ampli- tude-frequency response, as given by (40), (42) and (43), for Ω = 0 and Ω = 50, respectively. The other parameters are fixed as follows: α = 0.01, β = 0.05, γ = 0.1, h = 1 and a = 0.02. The solid line denotes the stable branch and the dashed line denotes the unstable one. For validation, analytical approxima- tions are compared to numerical integration (circles) using a Runge-Kutta method. The effect of Ω on the amplitude-frequency response is illustrated in Fig.

6(c)

for different values of Ω. The plots in this figure in- dicate that as Ω increases, the backbone curve shifts left and changes its bending from softening to hard- ening. Near the threshold Ω

c

given by (34), the back- bone curve becomes smaller, the frequency locking and hysteresis are reduced and jumps are attenuated.

Hysteresis is completely eliminated at the threshold Ω

c

corresponding to the vanishing of the nonlinear characteristic stiffness of the slow dynamics (11).

In Fig.

7, we show the

z-amplitude-frequency re-

sponse near the two resonances, 1:1 and 3:1. This fig-

ure indicates that for Ω = 0, the 3:1 resonance area is

in a narrow range comparing to the 1:1 resonance area

(Fig.

7(a)). In contrast, for

Ω = 50 the 3:1 resonance

curve becomes larger (Fig.

7(b)) indicating that in the

softening case the resonance 3:1 can be considered as

negligible, whereas in the hardening case the 3:1 reso-

nance is significant in terms of magnitude and width.

(9)

Fig. 6 Amplitude-frequency response of the slow dynamicsz(t ). Analytical approximation: solid (for stable) and dashed (for unsta- ble). Numerical simulation: circles. (a)Ω=0, (b)Ω=50, (c) effect of different values ofΩon the backbone curve

Fig. 7

Amplitude-frequency response near 1:1 and 3:1 resonances

4.2 Quasiperiodic modulation and hysteresis suppression

Following the same analysis as for the fundamental resonance case, we construct analytical approximation of the limit cycle of the slow flow (41) correspond- ing to quasiperiodic motion of the slow dynamics (11) near the 3:1 subharmonic resonance. The Cartesian system corresponding to the polar form (41) is writ- ten as

du

dt = Sv + η Au(Bu + Cv)

u

2

+ v

2

+ 2H

1

uvH

2

u

2

v

2

, dv

dt = − Su + η Av(BvCu)

u

2

+ v

2

+ H

1

u

2

v

2

+ 2H

2

uv ,

(44)

where the bookkeeping parameter η is introduced in damping and nonlinearity terms. Using the multiple scales method, the solution to the first order of system (44) is given by

u

0

(T

0

, T

1

) = R(T

1

) cos

ST

0

+ ϕ(T

1

) , v

0

(T

0

, T

1

) = − R(T

1

) sin

ST

0

+ ϕ(T

1

)

. (45)

The amplitude R and the phase ϕ vary with time ac- cording to the following slow-flow system:

dR

dt = ARBR

3

,

dt = −CR

2

.

(46)

(10)

Fig. 8 Entrainment area (E), hysteresis (H) and modulation amplitude of slow-flow limit cycle (QP)

The first-order approximate periodic solution of the slow flow (44) is then given by

u(t ) = R cos νt,

v(t ) = − R sin νt, (47)

where the amplitude R and the frequency ν are given by

R = A

B , (48)

ν = SCR

2

, (49)

and finally, the approximate quasiperiodic response of the slow dynamics (11) reads

z(t ) = R cos ω

3 + ν

t + F cos ωt. (50) In Fig.

8(a), (b), we show the analytical entrainment

zone along with the numerical quasiperiodic modula- tion domain of the slow-flow limit cycle for Ω = 0 and Ω = 50, respectively. The numerical quasiperi- odic modulation domain is marked with double circles connected with a vertical line.

Figure

8(a) shows the existence of two hysteresis

loops near this subharmonic resonance. One hystere- sis loop (right) is found to be very small, whereas the other one (left) is quite large and may produce a signif- icant jump phenomenon from higher to lower ampli- tude when the forcing frequency is swept backward.

In the hardening case (Fig.

8(b)), only one similar

jump phenomenon occurs when the forcing frequency is swept forward.

In Fig.

9, we present examples of time histories of

the slow dynamics z(t ) obtained by numerical simu- lation for some values of ω picked from Fig.

8(a). As

predicted, away from the resonance, a slow-flow limit cycle exists and attracts all initial conditions. The re- lated time trace is shown in subfigure (a) correspond- ing to ω = 2.72. In subfigures (b) and (c) we plot the time traces corresponding to ω = 2.8. These figures indicate the coexistence of periodic (stable slow-flow equilibrium) and quasiperiodic (stable slow-flow limit cycle) responses. As the forcing frequency increases, a transition from quasiperiodic vibration to a response at the 3:1 subharmonic frequency occurs. The region be- tween ω = 2.862 and ω = 2.939 is associated with this 3:1 frequency-locked motion, in which the response of the system follows the 3:1 subharmonic frequency.

The corresponding time trace is shown in subfigure (d) corresponding to ω = 2.9. It is worth noting that in the region corresponding to the small hysteresis, coexis- tence phenomenon can also occur. Moreover, saddle- node and heteroclinic bifurcations involving the slow- flow limit cycle and the cycles of order 3 can take place in this small region. However, this problem is beyond the scope of the current paper.

In Fig.

10, we show the variation of hysteresis area

(obtained from (42)) as a function of the frequency Ω for h = 1. This plot indicates that in the softening re- gion (Ω < Ω

c

) the hysteresis area is relatively small, whereas in the hardening region (Ω > Ω

c

) it becomes large. It can be seen from this figure that the hysteresis is attenuated around the threshold Ω

c

and completely eliminated at this threshold.

In Fig.

11, we plot the effect of different system

parameters on the attenuation domain of hysteresis.

Figure

11(a), (b) shows that varying the amplitude of

(11)

Fig. 9 Examples of time histories of the slow dynamicsz(t )

Fig. 10 Variation of the hysteresis area versus the frequencyΩ near the 3:1 subharmonic resonance

the external forcing h and the linear damping α has small effect on the attenuation area of hysteresis. Fig- ure

11(c) indicates that increasing the nonlinear damp-

ing β , the hysteresis area decreases and the attenuation

zone of hysteresis increases. Figure

11(d) illustrates

the effect of the amplitude a of the FHE on the hys- teresis. As it was seen in the primary resonance case, the plots in subfigure (d) indicate that increasing the amplitude a causes the attenuation domain of hystere- sis to shift left toward smaller values of Ω .

5 Conclusion

In this paper, we have investigated the effect of an FHE on hysteresis and on frequency locking in a forced van der Pol–Duffing oscillator near the primary and the subharmonic 3:1 resonances. Averaging procedure and multiple scales techniques are performed near each resonance to derive a slow dynamics and its slow flow. Analysis of equilibria of each slow flow provides analytical expressions of the corresponding backbone curve. The results showed that adding the FHE causes the backbone curves to shift left and the nonlinear characteristic to change from softening to hardening.

Analysis of the influence of different system parame-

ters on the hysteresis near the resonances reveals that

(12)

Fig. 11 (Color online) Effect of system parameters on frequency area.

(a) Effect of the external excitationh, (b) effect of the linear dampingα, (c) effect of the nonlinear dampingβ, (d) effect of the amplitudeaof the FHE

increasing the amplitude a of the FHE decreases the suppression domain and shifts it toward lower values of the fast frequency Ω . Moreover, adding the non- linear damping β increases the hysteresis suppression area.

Concerning the 3:1 subharmonic resonance, it is shown that in the softening situation (Ω < Ω

c

), the 3:1 resonance may occur in a small range of the forcing frequency, whereas in the hardening situation (Ω > Ω

c

), the 3:1 resonance domain is large (Fig.

7).

In addition, near this subharmonic resonance, two hys- teresis loops may exist, one of them is large (Fig.

8)

and can produce a significant jump that cannot be ig- nored in practical applications.

The results suggest that the high frequency Ω can be chosen so as to completely eliminate the hystere- sis loop for a significant range of the frequency near the 1:1 resonance and in a small range near the 3:1 resonance. This suppression offers a smooth transition between the quasiperiodic responses and the entrained motions, preventing jumps near the resonances.

References

1. Nayfeh, A.H., Younis, M.I.: Dynamics of MEMS res- onators under superharmonic and subharmonic excitations.

Micromech. Microeng. 15, 1840–1847 (2005)

2. Nayfeh, A.H., Younis, M.I., Abdel-Rahman, E.M.: Dy- namic pull-in phenomenon in MEMS resonators. Non- linear Dyn. 48, 153–163 (2007)

3. Pandey, M., Rand, R.H., Zehnder, A.T.: Perturbation analy- sis of entrainment in a micromechanical limit cycle oscil- lator. Commun. Non-linear Sci. Numer. Simul. 12, 1291–

1301 (2007)

4. Rajesh, S., Nandakumaran, V.M.: Control of bistability in a directly modulated semiconductor laser using delayed op- toelectronic feedback. Physica D 213, 113–120 (2006) 5. Fahsi, A., Belhaq, M., Lakrad, F.: Suppression of hys-

teresis in a forced van der Pol–Duffing oscillator. Com- mun. Non-linear Sci. Numer. Simul. (2008). doi:10.1016/

j.cnsns.2008.03.003

6. Stephenson, A.: On induced stability. Philos. Mag. 15, 233–

236 (1908)

7. Hirsch, P.: Das Pendel mit oszillierendem Aufhängepunkt.

Z. Angew. Math. Mech. 10, 41–52 (1930)

8. Kapitza, P.L.: Dynamic stability of a pendulum with an os- cillating point of suspension. Zurnal Eksp. Teor. Fiz. 21, 588–597 (1951). (In Russian)

(13)

9. Thomsen, J.J.: Some general effects of strong high- frequency excitation: stiffening, biasing, and smoothening.

J. Sound Vib. 253(4), 807–831 (2002)

10. Jensen, J.S., Tcherniak, D.M., Thomsen, J.J.: Stiffening ef- fects of high-frequency excitation: experiments for an axi- ally loaded beam. J. Appl. Mech. 67, 397–402 (2000) 11. Hansen, M.H.: Effect of high-frequency excitation on nat-

ural frequencies of spinning discs. J. Sound Vib. 234(4), 577–589 (2000)

12. Tcherniak, D., Thomsen, J.J.: Slow effect of fast harmonic excitation for elastic structures. Non-linear Dyn. 17, 227–

246 (1998)

13. Mann, B.P., Koplow, M.A.: Symmetry breaking bifurca- tions of a parametrically excited pendulum. Non-linear Dyn. 46, 427–437 (2006)

14. Sah, S.M., Belhaq, M.: Effect of vertical high-frequency parametric excitation on self-excited motion in a delayed van der Pol oscillator. Chaos Solitons Fractals 37(5), 1489–

1496 (2008)

15. Belhaq, M., Sah, S.M.: Horizontal fast excitation in delayed van der Pol oscillator. Commun. Non-linear Sci. Numer.

Simul. 13(8), 1706–1713 (2008)

16. Belhaq, M., Sah, S.M.: Fast parametrically excited van der Pol oscillator with time delay state feedback. Int. J. Non- linear Mech. 43(2), 124–130 (2008)

17. Belhaq, M., Fahsi, A.: 2:1 and 1:1 frequency-locking in fast excited van der Pol–Mathieu–Duffing oscillator. Non-linear Dyn. 53, 139–152 (2008)

18. Blekhman, I.I.: Vibrational Mechanics—Non-linear Dy- namic Effects, General Approach, Application. World Sci- entific, Singapore (2000)

19. Belhaq, M., Houssni, M.: Quasi-periodic oscillations, chaos and suppression of chaos in a nonlinear oscillator driven by parametric and external excitations. Non-linear Dyn. 18, 1–24 (1999)

20. Belhaq, M., Guennoun, K., Houssni, M.: Asymptotic solu- tions for a damped non-linear quasi-periodic Mathieu equa- tion. Int. J. Non-linear Mech. 37, 445–460 (2000) 21. Rand, R.H., Guennoun, K., Belhaq, M.: 2:2:1 resonance in

the quasi-periodic Mathieu equation. Non-linear Dyn. 31, 187–193 (2003)

22. Rand, R.H., Morrison, T.: 2:1:1 resonance in the quasi- periodic Mathieu equation. Non-linear Dyn. 40, 195–203 (2005)

23. Sah, S.M., Recktenwald, G., Rand, R.H., Belhaq, M.: Au- toparametric quasiperiodic excitation. Int. J. Non-linear Mech. 43, 320–327 (2008)

24. Nayfeh, A.H., Mook, D.T.: Non-linear Oscillations. Wiley, New York (1979)

Références

Documents relatifs

A perturbation technique is then performed on the slow dynamic near the 2:1 and 1:1 resonances, re- spectively, to obtain reduced autonomous slow flow equations governing the

where r max is the upper limit of the modulation ampli- tude of the slow flow limit cycle given by (23) (first- order approximation) or by (31) (second-order approx- imation) and r

The results sug- gest that in some practical situations in which the delay is imposed and the fast excitation is induced by an exter- nal vibrational environment, the tilt angle of

More clearly, it can be concluded that the contact stiffness characteristic of the sample, the shift in the contact resonance frequency and the interval of the unstable trivial

In this paper, we study the influence of adding fast harmonic (FH) excitation on the entrainment area of the main parametric resonance for a self- and parametrically (SP)

To analyse the effect of vertical FH parametric excitation and time delay on the suppression of limit cycle in van der Pol oscillator, we apply, in a first step, the method of

To analyse the effect of vertical FH parametric exci- tation and time delay state feedback on the suppres- sion of limit cycle in van der Pol oscillator, we apply, in a first step,

It is shown that this analytical method, based on a double averaging procedure, is efficient to capture the modulation domain of the quasiperiodic response as well as the threshold