You are on page 1of 9

Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001

www.elsevier.com/locate/nahs

Chaos control in AFM systems using nonlinear delayed feedback via


sliding mode control
Mehdi Tabe Arjmand, Hoda Sadeghian, Hassan Salarieh, Aria Alasty ∗
Center of Excellence in Design, Robotics and Automation, Department of Mechanical Engineering, Sharif University of Technology,
PO Box 11365-9567, Tehran, Iran

Received 25 May 2007; accepted 11 October 2007

Abstract

In this paper a nonlinear delayed feedback control is proposed to control chaos in an Atomic Force Microscope (AFM) system.
The chaotic behavior of the system is suppressed by stabilizing one of its first-order Unstable Periodic Orbits (UPOs). At first, it
is assumed that the system parameters are known, and a nonlinear delayed feedback control is designed to stabilize the UPO of
the system. Then, in the presence of model parameter uncertainties, the proposed delayed feedback control law is modified via
sliding mode scheme. The effectiveness of the presented methods is numerically investigated by stabilizing the unstable first-order
periodic orbit of the AFM system. Simulation results show the high performance of the methods for chaos elimination in AFM
systems.
c 2007 Elsevier Ltd. All rights reserved.

Keywords: AFM; Chaos; Delayed feedback; Sliding mode

1. Introduction

Atomic force microscopy (AFM) has been widely used for surface inspection with nanometer resolution in
engineering applications and fundamental research since the time of its invention, 1986 [1]. As an imaging tool,
the AFM is capable of resolving surface features at the atomic level for conducting and non-conducting samples.
Currently, the AFM is used in many imaging applications ranging from biological systems to semiconductor
manufacturing.
The mechanism of AFM basically depends on the interaction of a microcantilever with surface forces. The tip
of the microcantilever interacts with surface through a surface–tip interaction potential. One approach to measure
the surface forces is to monitor the deflection of the microcantilever through a photodiode. This approach is named
“contact mode”. Another approach termed “tapping mode”, is performed by vibrating the microcantilever close to its
resonance frequency and monitoring the changes in its effective spring constant. In this method, the driving amplitude
is set to a constant value and typical resonant frequencies are in the range from a few kilohertz to some megahertz [2].
A microcantilever in tapping mode may exhibit chaotic behavior under certain conditions. This matter has been
experimentally observed in [3–5]. Theoretical studies, based on the techniques of Melnikov theory, have been

∗ Corresponding author. Tel.: +98 21 6616 5504; fax: +98 21 6600 0021.
E-mail address: aalasti@sharif.edu (A. Alasty).

c 2007 Elsevier Ltd. All rights reserved.


1751-570X/$ - see front matter
doi:10.1016/j.nahs.2007.10.002
994 M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001

performed in [1,6] to prove the existence of chaotic invariant sets and to determine the region in the space of physical
parameters where chaotic motion is present. In addition, using nonlinear analysis techniques on attracting limit sets,
the the presence of chaotic invariant sets is numerically verified in [2,7].
Typically, in theoretical and numerical works, the interaction between the sample and the microcantilever is
modeled by a Lennard-Jones potential which consists of a short-range repulsive potential and a long-range Van
der Waals attractive potential. By analyzing the dynamics of the microcantilever–sample system subjected to a
harmonic force, it is concluded that chaos can exist in AFM depending on the damping value, excitation amplitude
and frequency, and average tip–sample distance.
The existence of chaotic behavior in AFM is highly undesirable for its performance since this type of complex
irregular motion causes the AFM to give inaccurate measurements and low resolution of the achieved sample
topography. Accordingly, it is always required to ensure good performance of the microscope and to eliminate the
possibility of chaotic motion of the microcantilever either by changing the AFM operating conditions to a region
of the parameter space where regular motion is assured [8,9] or by designing an active controller that stabilizes the
system on one of its UPOs [10].
Chaos control via stabilizing a UPO has attracted a great deal of research interest in recent years. In this area, some
techniques have been developed based on low energy control signals [11,12]. Also, Pyrages designed a controller
based on the linear delayed feedback control [13] that has been extensively employed for stabilizing periodic solutions
of nonlinear systems [14]. Other approaches like sliding mode, Fuzzy control and feedback linearization also have
been used in diverse chaotic systems [15–17].
In this paper, we have developed a nonlinear delayed feedback control algorithm to stabilize the AFM
microcantilever on its first-order UPO when specific parameters lead the system into a chaotic zone. This UPO is
actually an unstable solution of the AFM system with a period equal to that of the exciting force on the system.
Also a delayed feedback control via sliding mode scheme is proposed for chaos elimination in the presence of some
uncertainties in the model parameters. The methods are used for stabilizing the first-order periodic orbits of a chaotic
AFM to numerically investigate the effectiveness of the control algorithm. Simulation results show that the proposed
techniques can be successfully implemented for chaos suppression in the AFM, even when the exact values of model
parameters are not accessible. Moreover, the robustness of the designed controller is checked for possible noise in the
measured states.

2. Problem statement

The forced dynamical system describing tapping mode operation can be obtained based on the model suggested
by Ashhab et al. in [1]. In their work, the microcantilever tip–sample interaction is modeled by the Lennard-Jones
interaction potential. The system can be written in a non-dimensionalized form as [7]:

ξ̇1 (t) = ξ2 (t)



(1)
ξ̇2 (t) = −δξ2 (t) − ξ1 (t) + f˜ cos(Ω t) + f IL (ξ1 (t)),
where t is the non-dimensional time; ξ1 (t) and ξ2 (t) are the non-dimensional position and velocity of the
microcantilever tip; f˜ and Ω are the amplitude and the frequency of the forcing term, respectively; δ is the damping
factor; and

σ̃ 6 d d
f IL (ξ1 (t)) = − (2)
30 (α + ξ1 (t))8
(α + ξ1 (t))2
denotes the attraction/repulsion interaction force derived from Lennard-Jones interaction potential.
Under certain conditions for the values of σ̃ , d, α, Ω , δ and f˜, the above described system exhibits chaotic behavior.
The main goal is to stabilize the system (1) on one of its UPOs when chaotic motion emerges. Here, the unstable
periodic orbit of the system with a period equal to that of the exciting force, i.e. 2π
Ω , is chosen as the desired trajectory
for the system to follow. To this effect, based on the assumption that both the state variables ξ1 (t) and ξ2 (t) are
available, at first a nonlinear delayed feedback control is proposed to impose the chaotic system to follow a regular
periodic trajectory. Then, a delayed feedback control via sliding mode scheme is proposed for the case that there exists
some uncertainties in the values of parameters in the microcantilever tip–sample model, e.g. d, σ̃ and α. In this case,
M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001 995

it is assumed that the uncertainties in the mentioned parameters are all bounded and the lower and upper bounds of
the uncertainties for each parameter are known. Note that in practice the parameter α is not exactly known but its
approximated value and its lower and upper limits can be easily obtained.

3. Chaos control of AFM

In the standard tapping mode operation, the microcantilever is sinusoidally forced by a piezo actuator near its
natural frequency, and periodically ‘taps’ the surface [10]. The implementation of nonlinear control techniques for
suppressing chaotic dynamics in the AFM system will be achieved by superimposing a feedback control forcing
signal onto the sinusoidal forcing signal both produced by the piezo. The new equations of motion including the
dimensionless form of control force, u, can be written as:
ξ̇1 (t) = ξ2 (t)

(3)
ξ̇2 (t) = −δξ2 (t) − ξ1 (t) + f˜ cos (Ω t) + f IL (ξ1 (t)) + u(t)
or equivalently:

ξ̈1 (t) = −δ ξ̇1 (t) − ξ1 (t) + f˜ cos (Ω t) + f IL (ξ1 (t)) + u(t). (4)
A proper control method to eliminate chaos via stabilizing a UPO in the AFM system is nonlinear delayed feedback
control. In the face of model parameter uncertainties, a modified version of this technique, named nonlinear delayed
feedback control via sliding mode, may be applied to the system. These two methods are described in the following
subsections.

3.1. Delayed feedback control via sliding mode

An approach to control chaos used by numerous authors is the so-called delayed feedback control which was
originally proposed by Pyragas. This method uses a linear combination of the error signals defined as:
ei (t) = ξi (t) − ξi (t − T ), (5)
where T is a delay time which can be equal to the period of a UPO of the system, and ξi (t) is a state variable.
In [18] a nonlinear delayed feedback control has been presented which modifies the commonly used Pyragas’ method
to achieve much better performance in chaos control. Here, at first a brief review of the method proposed in [18] is
presented.
Consider the delayed error in Eq. (5) as:
e(t) = ξ1 (t) − ξ1 (t − T ), (6)
where T = 2π Ω is the period of the exciting force in the AFM system of Eq. (4). Rewriting Eq. (4) for a T time units
earlier yields:

ξ̈1 (t − T ) = −δ ξ̇1 (t − T ) − ξ1 (t − T ) + f˜ cos (Ω t − Ω T ) + f IL (t − T ) + u(t − T ), (7)


where f IL (t) is f IL (ξ1 (t)). Subtracting Eq. (7) from (5) and using the error state defined in (6) one may write,
ë(t) = −δ ė(t) − e(t) − f IL (t − T ) − u(t − T ) + f IL (t) + u(t). (8)
We need to stabilize the error signal and enforce it to converge asymptotically to zero. To do so, the controlling signal
is proposed to be:
u(t) = u(t − T ) + f IL (t − T ) − f IL (t) + δ ė(t) + e(t) − [λ1 ė(t) + λ2 e(t)]. (9)
Thus, Eq. (9) would be,
ë(t) + λ1 ė(t) + λ2 e(t) = 0. (10)
λ1 and λ2 are selected such that the above error dynamics is exponentially stable. So the system with the designed
Ω ).
controller is asymptotically stable on its periodic solution (T = 2π
996 M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001

In practice some uncertainties may exist, especially on the α parameter, in Eq. (2). Here, we design a control to
overcome possible uncertainties in the system model; hence the control algorithm can be practically applied in actual
cases. For simplicity in formulations, we first define:

χ , σ̃ 6 d. (11)
The uncertainties are assumed to be on the values of d, χ and α as:

|α − ᾱ| < εα , d − d̄ < εd , |χ − χ̄ | < εχ ,



(12)

where ᾱ, d̄ and χ̄ are the nominal known or approximated values of α, d and χ . Also εα , εd and εχ are the known
upper bounds of the errors between the actual and nominal values of α, d and χ , respectively.
Now define:
ξ1 (t − T ) = ξ̄1
(13)
u(t − T ) = ū.
Adding and subtracting some equal terms, Eq. (8) is rewritten as:

ë + δ ė + e = u − ū + f IL (ξ̄1 + e) − f¯IL (ξ̄1 + e) + f¯IL (ξ̄1 + e) − f IL (ξ̄1 ) + f¯IL (ξ̄1 ) − f¯IL (ξ̄1 ), (14)
where,

χ̄ d̄
f¯IL (ξ̄1 ) = 8 − 2 . (15)
30 ᾱ + ξ̄1 ᾱ + ξ̄1
One can write,
χ d χ̄ d̄
f IL (ξ̄1 ) − f¯IL (ξ̄1 ) = 8 − 2 − 8 + 2
30 α + ξ̄1 α + ξ̄1 30 ᾱ + ξ̄1 ᾱ + ξ̄1
χ − χ̄ + χ̄ d − d̄ + d̄ χ̄ d̄
= 8 − 2 − 8 + 2 . (16)
30 α + ξ̄1 α + ξ̄1 30 ᾱ + ξ̄1 ᾱ + ξ̄1
Using Eq. (12), it is concluded that:
f IL (ξ̄1 ) − f¯IL (ξ̄1 ) < N (α, ξ̄1 ),

(17)
where,
1 1
N (α, ξ̄1 ) = εχ  8 + εd 2
30 α + ξ̄1 α + ξ̄1

1 1 1 1
+ χ̄ − + d̄ − . (18)

30 α + ξ̄ 8 8 2 2

30 ᾱ + ξ̄1 α + ξ̄1 ᾱ + ξ̄1
   
1

Now considering the variation of α according to Eq. (12), it is obvious that


f IL (ξ̄1 ) − f¯IL (ξ̄1 ) < max(N (ᾱ − εα , ξ̄1 ), N (ᾱ + εα , ξ̄1 )) , M(ξ̄1 ).

(19)
Similarly it can be obtained that
f IL (ξ̄1 + e) − f¯I L (ξ̄1 + e) < max(N (ᾱ − εα , ξ̄1 + e), N (ᾱ + εα , ξ̄1 + e)) = M(ξ̄1 + e).

(20)
Therefore defining,

η(εα , εd , εχ , ξ̄1 , e) , f IL (ξ̄1 + e) + f¯IL (ξ̄1 + e) − f IL (ξ̄1 ) − f¯IL (ξ̄1 ) (21)


M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001 997

and using the triangular inequality, it is concluded that


η(εα , εd , εχ , ξ̄1 , e) ≤ M(ξ̄1 + e) + M(ξ̄1 ).

(22)
So one can rewrite Eq. (14) as:

ë + δ ė + e = u − ū + f¯IL (ξ̄1 + e) − f¯IL (ξ̄1 ) + η(εα , εd , εχ , ξ̄1 , e), (23)


where η has an uncertainty with the bound introduced in Eq. (22). Now let

u − ū = v − f¯IL (ξ̄1 + e) + f¯IL (ξ̄1 ), (24)


where v is a part of controlling signal which will be obtained during this section. Substituting Eq. (24) into Eq. (23)
yields:

ë + δ ė + e = v + η(εα , εd , εχ , ξ̄1 , e). (25)


To design a control law which asymptotically stabilizes the origin of the above error dynamics, we use the sliding
mode method by defining a sliding surface as,
S = ė + λe, (26)
where λ is a positive constant. A Lyapunov function is defined as:
1 2
V = S . (27)
2
The derivative of V along the error trajectories is:

V̇ = S Ṡ = S(ë + λė) = S(−δ ė − e + v + η + λė). (28)


Set v as:
v = δ ė + e − λė − K (t)sign(S), (29)
where K (t) must satisfy the condition of

K (t) ≥ M(ξ̄1 + e) + M(ξ̄1 ) + θ (30)


and θ is a positive constant. Substituting Eqs. (29) and (30) into Eq. (28) results in:

V̇ = S(η − K (t)sgn(S)) < −θ |S| . (31)


It is concluded that the sliding surface S = 0 is attracted by the error trajectories in a finite time, hence the tracking
error asymptotically converges to zero. Thus considering Eq. (29) the control law can be summarized as:

u(t) = u(t − T ) − f¯IL (ξ̄1 + e) + f¯IL (ξ̄1 ) + δ ė + e − λė − K (t)sign(S). (32)

3.2. Control implementation remarks

Although the controlled system on the sliding surface describes a stable system, the value of corresponding u in Eq.
(32) may vary in a wide range due to large value of nonlinear attraction/repulsion force. As the control signal cannot
exceed the limited force range of the piezo, we need a bounded control signal when the proposed control method is
experimentally implemented.
Suppose that the control signal is limited to be in the range of [−û, û]. Therefore Eq. (32) is rewritten as:
u 0 (t) = u(t − T ) + fˆIL (t − T ) − fˆIL (t) + δ ė(t) + e(t) − [λ1 ė(t) + λ2 e(t)]
u(t) = u 0 (t) |u 0 (t)| < û

(33)
u(t) = û · sign(u 0 (t)) |u 0 (t)| > û.
998 M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001

Fig. 1. Chaotic attractor of the AFM (a) in the phase space, (b) in the Poincare section (10 000 points).

Moreover, when the error signal reaches a desired magnitude, i.e. the UPO of the system is being stabilized, the
saturation bounds can be decreased because when the UPO of the uncontrolled system is achieved, the control should
be equal to zero. For example one may use an exponential function for this purpose:

û(t) = as ebs (t−tˆ) . (34)


Here as and bs are two arbitrary constants, and tˆ is the time when the error signal is sufficiently close to zero.
Decreasing the saturation bound guaranties that the stabilized periodic solution is a UPO of the uncontrolled
system. It must be noted that in recursive control techniques such as the method proposed in this paper and the
one presented in [18], the stabilized periodic solution may not be the UPO of the uncontrolled system. The UPO of
the uncontrolled system is obtained when u(t) converges to zero, so by decreasing the saturation limit this condition
is satisfied, nevertheless setting any saturation bound may have undesired effects on the stabilization property of the
controller. The saturation bound is chosen arbitrarily since it depends on the real-world condition not the controller
algorithm.
There are some other methods which guaranty that the stabilized periodic orbit is the UPO of the uncontrolled
system. One of them is to apply the controller when the delayed error becomes smaller than a sufficiently small desired
value. This condition will be satisfied because chaotic systems have the recurrence or ergodic property. However all
of these techniques may result in decreasing the controller performance.

4. Simulation results

The above described control scheme is now used to control the states of a chaotic AFM system. The aforementioned
AFM system in Eq. (1) exhibits chaotic behavior with the following parameters [2]:
4
δ = 0.04, σ̃ = 0.3, α = 0.8, f˜ = 2.0, Ω = 1, . d= (35)
27
For these parameters, the phase plane diagram of the system chaotic attractor is shown in Fig. 1(a). Also this attractor,
in the Poincare section of t = nT (T = 2π Ω ), is illustrated in Fig. 1(b).
Now the objective is to eliminate the chaotic behavior by stabilizing the system on its one cycle UPO, i.e. with the
period of T = 2πΩ , using the delayed feedback control scheme described in Section 3.
The simulation consists of two parts. First in the absence of uncertainties, the described controller in [18] is
implemented. Then, by considering some uncertainties in the model parameters, the proposed delayed feedback via
sliding mode is applied to the system. In both the methods the controller signal is equal to zero for the first T time
units. For time greater than T , the controller acts in a constant saturation limit until the stability of the periodic orbit
is assured. This matter can be checked by observing the position error, e1 (t). The constant saturation bound in the
simulation is set to 0.1. When the position error becomes less than a desired magnitude, that we have set to 0.1, the
exponential saturation in Eq. (34) with as = 0.1 and bs = −0.1 would be the saturation limit of the control signal.
This guaranties that the stabilized periodic orbit is a UPO of the uncontrolled AFM.
M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001 999

Fig. 2. Simulation results for the nonlinear delayed feedback control (a) ξ1 time series (b) u time series (c) Stabilized UPO.

The controller in delayed nonlinear feedback control follows the pattern defined in Eq. (9), with λ1 = 3 and λ2 = 2.
Fig. 2 shows the results for this case.
In the presence of uncertainties in the system, the sliding mode delayed feedback is used to control the chaotic
motion of the system. Regarding Eq. (9), the following uncertainties are considered in the simulation:

εα = 0.1ᾱ εd = 0.1d̄ εχ = 0.1χ̄ , (36)

where ᾱ = 0.3, d̄ = 27 4
and χ̄ = 0.000108. Also, in Eq. (36) λ and θ are set to 1.
Fig. 3 represents the results of the sliding mode version of the controller. Also, in this case, the robustness of the
designed controller against the output noise is checked for 1% noise added to the measured values of states. The
results are shown in Fig. 4.
As can be seen from the simulation results, the stability of unstable periodic orbits is completely achieved in less
than 100 non-dimensional time units, which is in the order of 0.01 s in a real AFM system. The amplitude of the
control signal is sufficiently small and converges to zero in a finite small duration of time, and the system is strongly
robust to disturbances and measurement noise.
To show the ability of the controller to stabilize periodic solutions in the absence of saturation, an extra simulation
is done by setting the bound of saturation to infinity. The results are shown in Fig. 5. It is observed that the rate
of convergence towards the periodic orbit is higher than the previous examples. Besides the control action does not
vanish, because the stabilized orbit is nearly different from the main UPO of the uncontrolled system.

5. Conclusion

In this paper, a new method is developed to eliminate chaos in AFM system based on stabilizing a UPO of the
system in the presence of model uncertainties. The nonlinear delayed feedback control is developed and used to control
1000 M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001

Fig. 3. Simulation results for the sliding mode control (a) ξ1 time series (b) u time series (c) Stabilized UPO.

Fig. 4. Simulation results for the nonlinear delayed feedback control via sliding mode with a 1% noise in feedback states (a) ξ1 time series (b) u
time series.
M.T. Arjmand et al. / Nonlinear Analysis: Hybrid Systems 2 (2008) 993–1001 1001

Fig. 5. Simulation results for the nonlinear delayed feedback control via sliding mode with unsaturated control action (a) ξ1 time series (b) u time
series.

the chaotic AFM system when it is assumed that the exact values of model parameters are available. The modified
version, the nonlinear delayed feedback via sliding mode, is applied to the system by considering some uncertainties
in the parameters. Also the robustness of the system is checked for possible measurement noise in the feedback states.
Both the techniques are successfully implemented to stabilize the period-1 UPO of the AFM system. Simulation
results confirm the validity of the designed controllers. In both the methods, the system is properly stabilized on a
periodic trajectory in a reasonable small time and control action.

References
[1] M. Ashhab, M.V. Salapaka, M. Dahleh, I. Mezic, Dynamical analysis and control of micro-cantilevers, Automatica 35 (1999) 1663–1670.
[2] M. Basso, L. Giarrk, M. Dahleh, I. Mezic, Numerical analysis of complex dynamics in atomic force microscopes, Proc. IEEE Int. Conf.
Control Appl. (1998).
[3] N.A. Burnham, A.J. Kulik, G. Gremaud, G.A.D. Briggs, Nanosubharmonics: The dynamics of small nonlinear contacts, Phys. Rev. Lett.
74 (1995) 5092–5059.
[4] F. Jamitzky, M. Stark, W. Bunk, W.M. Heckl, R.W. Stark, Chaos in dynamic atomic force microscopy, Nanotechnology 17 (2006) 213–220.
[5] Daniel S. Burgess, Chaos affects atomic force microscopes, Photon. Spectra 40 (4) (2006) 101–102.
[6] M. Ashhab, M.V. Salapaka, M. Dahleh, I. Mezic, Melnikov-based dynamical analysis of microcantilevers in scanning probe microscopy,
Nonlinear Dynam. 20 (3) (1999) 197–220.
[7] M. Basso, L. Giarrk, M. Dahleh, I. Mezic, Complex dynamics in a harmonically excited Lennard-Jones Oscillator: Micro-cantilever-sample
interaction in scanning probe microscopes, J. Dyn. Syst. Meas. Control 122 (2000) 240–245.
[8] M. Ashhab, M. Salapaka, Rail Dahleh, I. Mezic, Control of Chaos in Atomic Force Microscopes, ACC, Albuquerque, New Mexico, 1997.
[9] M. Basso, G. Bagni, Controller synthesis for stabilizing oscillations in tapping-mode atomic force microscopes, IEEE Inter. Symp. Comput.
Aided Control Sys. Design (2004) 372–377.
[10] M. Stonier, Dynamical analysis and control of nonlinear instabilities in nano-devices, Thesis for BE, The University of Queensland, 2002.
[11] E. Ott, C. Grebogi, J.A. York, Controlling chaos, Phys. Rev. Lett. 64 (1990) 1196–1199.
[12] T. Shinbort, E. Ott, C. Grebogi, J. Yorke, Using chaos to direct trajectories to targets, Phys. Rev. Lett. 65 (1990) 3215–3218.
[13] K. Pyragas, Continuous control of chaos by self-controlling feedback, Phys. Lett. A. 170 (1992) 421–428.
[14] G. Chen, X. Yu, On time delayed feedback control of chaotic systems, IEEE Trans. Circuits Syst. 1 46 (1999) 767–772.
[15] H. Layeghi, M.T. Arjmand, H. Salarieh, A. Alasty, Stabilizing periodic orbits of chaotic systems using fuzzy adaptive sliding mode control,
Chaos Solitons Fractals (2006), doi:10.1016/j.chaos.2006.10.021.
[16] M. Bonakdar, M. Samadi, H. Salarieh, A. Alasty, Stabilizing unstable periodic orbits of chaotic systems using fuzzy control of Poincare map,
Chaos Solitons Fractals (2006), doi:10.1016/j.chaos.2006.06.081.
[17] H. Salarieh, M. Shahrokhi, Indirect adaptive control of discrete chaotic systems, Chaos Solitons Fractals 34 (2007) 1188–1201.
[18] J. Zhu, Y.-P. Tiang, Stabilizing periodic solutions of nonlinear systems and applications in chaos control, IEEE Trans. Circuits Syst. II 52 (12)
(2005) 870–874.

You might also like