Chapter 6

Chaos is observed as an unpredictable phenomenon due to its sensitivity to initial states. It is a kind of steady-state but locally unstable behavior, and exhibits irregular properties. Such random-like phenomenon is normally regarded as unstable operation which results in additional loss, and therefore is a harmful behavior. Various control methods have been proposed to stabilize the chaotic behavior, such as the Ott–Grebogi–Yorke (OGY) method (Ott, Grebogi, and Yorke, 1990; Hunt, 1991), the time-delay feedback method (Pyragas, 1992), the non-feedback method (Rajasekar, Murali, and Lakshmanan, 1997; Ramesh and Narayanan, 1999), the proportional feedback method (Jackson and Grosu, 1995; Casas and Grebogi, 1997), the nonlinear control method (Khovanov et al., 2000; Tian, 1999), the adaptive control method (Boccaletti, Farini, and Arecchi, 1997; Liao and Lin, 1999), the neutral networks method (Hirasawa et al., 2000; Poznyak, Yu, and Sanchez, 1999), and the fuzzy control method (Tanaka, Ikeda, and Wang, 1998). Some of them have also been proposed to stabilize the chaotic behavior in electric drive systems.

In this chapter, various control approaches, including the time-delay feedback control, the nonlinear feedback control, the backstepping control, the dynamic surface control and the sliding mode control, are introduced to stabilize the chaos that occurs in both DC and AC drive systems.

6.1 Stabilization of Chaos in DC Drive System

6.1.1 Modeling

As shown in Figure 6.1, the time-delay feedback control method is used to stabilize chaos in a voltage-controlled DC drive system. Time-delay feedback control has some definite advantages: it does not desire a priori analytical knowledge of the system dynamics; it does not require a reference signal corresponding to the desired unstable periodic orbit (UPO); and it does not need fast sampling or a computer analysis of the state of the system. Also, the corresponding perturbation is small when the delayed time is close to the period of the desired UPO (Kittel, Parisi, and Pyragas, 1995).

Figure 6.1 Voltage-controlled DC drive system with time-delay feedback control

img

The speed control of the DC drive is implemented by constant-frequency pulse width modulation (PWM). The ramp voltage signal img, which functions to generate the PWM signal, is represented by:

(6.1) equation

where img, img, and T are the lower limit, upper limit and period of the sawtooth wave. On the other hand, the speed error signal img is given by:

(6.2) equation

where g is the speed feedback gain, which is actually the overall gain of the speed encoder, frequency-to-voltage (F/V) converter, and operational amplifier 1 (OA1), ω is the actual speed, and img is the reference speed. This speed error is chosen as the feedback variable. By using the bucket-brigade delay (BBD) line, the corresponding delayed signal img for the time-delay feedback control is described as:

(6.3) equation

where τ is the time delay. When the desired orbit is the embedded unstable period-p orbit, img is normally chosen to be pT. Thus, the feedback control signal img can be represented as:

(6.4) equation

where η is the delayed feedback gain, which is actually the gain in the operational amplifier 2 (OA2). The operational amplifier 3 (OA3) simply adopts a unity gain. Then, img and img are fed to the comparator (CM), and thus generate the PWM signal to turn on and off the power switch S. When img is larger than img, S is turned off and the diode D conducts. Otherwise, S is turned on and D is turned off. The inductor L is connected in series with the DC motor to ensure a continuous conduction mode of operation. So, the voltage-controlled DC drive system with a time-delay feedback control can be modeled as:

(6.5) equation

It should be noted that img comprises a delayed component, and the dynamical equation in (6.5) is actually a time-delayed differential equation. The delayed feedback only exists in the switching condition img. As discussed in Chapter 3, the corresponding Poincaré map with multiple switching pulses can be represented by:

(6.6) equation

The switching points img can be determined by using the switching condition described in (6.5):

(6.7) equation

Then, the switching points between the interval img and img are given by:

(6.8) equation

The corresponding system states in (6.7) are given by:

(6.9) equation

Thus, the switching points can be calculated by using (6.6), (6.7), and (6.9), which yields:

(6.10) equation

Given the values of img, the values of img can be calculated by using (6.10).

For a practical voltage-controlled DC drive system, there is only a single switching pulse. So, the corresponding Poincaré map can be represented by:

(6.11) equation

When img, the instant of switching action in (6.10) is given by:

(6.12) equation

When img, the instant of switching action in (6.10) is given by:

(6.13) equation

Since img can be represented by img and img, (6.13) is rewritten as:

(6.14) equation

It should be noted that img is the switching conditions when img is on the left side of img, and img is the switching conditions when img is on the right side of img. The stability of control can be described by (6.11), (6.12), and (6.14). Since τ is normally chosen to be the period T, namely p = 1, the corresponding Poincaré map can be simplified as:

(6.15) equation

(6.16) equation

(6.17) equation

When img, img is an explicit function of img; and when img, img is an implicit function of img and img. Let img, the Poincaré map can be rewritten as:

(6.18) equation

(6.19) equation

where img and img are the Poincaré maps corresponding to the conditions that img and img, respectively.

6.1.2 Analysis

The Poincaré map with the time-delay feedback control is different from that of the system without control. By choosing a suitable delay feedback gain η, the chaotic behavior can be stabilized into periodic behavior. The effective range for stabilization is called as the stable domain.

6.1.2.1 Fundamental Operation

The fundamental operation of the DC drive system corresponds to the period-1 orbit, namely img. So, the delay component in (6.16) and (6.17) become zero. Namely, img and img are independent of η. Therefore, the fixed point of period-1 orbit can be written as:

(6.20) equation

The characteristic multiplier of (6.20) is also independent of η. According to the implicit theorem, the Jacobian matrices of (6.18) and (6.19) are given by:

(6.21) equation

(6.22) equation

where

(6.23) equation

(6.24) equation

(6.25) equation

(6.26) equation

(6.27) equation

Since img, img can be represented by:

(6.28) equation

Theorem 6.1

When the eigenvalues of a matrix img are zero, the eigenvalues of the matrix img are also zero.

Proof

When img is the Jordan normalization form of img, it results in img. Thus, it yields img. When the eigenvalues of img are equal to zero, img can be described as img.

Hence, it deduces that img, and the eigenvalues of img are also zero.

Theorem 6.2

When the two eigenvalues of img in (6.25) are zero, two of the four eigenvalues of img are zero, and the other two eigenvalues of img are equal to those of img. These two eigenvalues are also equal to the eigenvalues of the matrix img, which is expressed as:

(6.29) equation

Proof

Given img and img, yields img.

From (6.23) and (6.24), it can be deduced that img. The eigenvalues of img are the same as those of img. Two of the eigenvalues of img are the eigenvalues of img, and the other two are the eigenvalues of img. Thus, from (6.23)(6.26), img can be described as:

(6.30) equation

According to Theorem 1, when the eigenvalues of img are all equal to zero, the eigenvalues of img are zero. When img and img, the eigenvalues of img are zero. From (6.23)(6.28), img can be described as:

(6.31) equation

So, the eigenvalues of img or img are equal to the eigenvalues of img in (6.29).

Because

img

the eigenvalues of S are equal to zero. Thus, the aforementioned theorems can be applied to a voltage-controlled DC drive system. The characteristic multipliers of a period-1 orbit can be calculated by the eigenvalues of img in (6.29). By calculating the characteristic multipliers, the stable domain of the control parameters – namely the time-delay feedback gain versus various system parameters for effective stabilization – can be determined. It should be noted that the calculation of the eigenvalues of img is easier than those for img and img. In particular, the dimension of img is twice that of img.

6.1.2.2 Subharmonic Operation

The stable domain for the period-p (p > 1) is illustrated by the period-2 operation. When img, a second-order Poincaré map of the drive system can be obtained by using (6.18) and (6.19):

(6.32) equation

(6.33) equation

Contrary to the fixed point of the period-1 orbit, the fixed point of the period-2 orbit is dependent on η, which can be represented by:

(6.34) equation

(6.35) equation

Substituting (6.34) into (6.35), img and img can be deduced by solving (6.35), and the period-2 orbit img can be obtained from (6.34). Similar to the period-1 orbit, the characteristic multipliers for the period-2 orbit are the eigenvalues of the Jacobian matrix img which can be deduced from the equation:

(6.36) equation

where

(6.37) equation

(6.38) equation

By calculating the characteristic multipliers, the corresponding stable domain of the control parameters can be obtained.

6.1.3 Simulation

In order to validate the aforementioned stabilization of chaos, a computer simulation is carried out. The parameters adopted for the simulation are based on a practical voltage-controlled DC drive system, namely R = 4.1 Ω, KE = 0.1356 V/rad/s, KT = 0.1324 Nm/A, J = 0.000557 kgm2, B = 0.000275 Nm/rad/s, L = 28 mH, Tl = 0.5 Nm, img = 2.2 V, vl = 0 V, T = 6.667 ms, and img. The delayed feedback gain η, speed feedback gain g, and supply voltage img are the control parameters for this simulation.

Based on the above stability analysis, the stable domain of η versus g under img = 60 V for period-1 and period-2 behaviors is depicted in Figure 6.2a, while the stable domain of η versus img under g = 1.4 V/rad/s is depicted in Figure 6.2b. From Figure 6.2a, it can be seen that when g < 0.8 V/rad/s the stable domain of period-1 motion D1 embraces the case η = 0. This means that when g < 0.8 V/rad/s, a DC drive system without a time-delay feedback control can operate stably. Similar to the period-1 orbit, a DC drive system without a time-delay feedback control can also exhibit a stable period-2 orbit within the domain D2 when 0.8 V/rad/s < g < 1.15 V/rad/s. It also can be observed that, with the increase of g and Vs, the range of η for D1 and D2 become narrower. This feature indicates that with an increase of g and Vs, the corresponding stable domain becomes narrower and chaos is more prone to occur. The same phenomena can be observed in Figure 6.2b.

Figure 6.2 Stable domains. (a) Delayed feedback gain versus speed feedback gain. (b) Delayed feedback gain versus supply voltage

img

img

In order to illustrate the stabilization of chaos, the trajectory of armature current i versus feedback control signal img is used for analysis. This feedback control signal is actually used to describe the motor speed. When η = 0, Vs = 60 V and g = 1.6 V/rad/s, the system operates in a chaotic mode, and the trajectory of i versus vc is as plotted in Figure 6.3a. It can be found that the trajectory exhibits a random-like but bounded behavior, with boundaries of img [0 V, 3.3 V] and img [1.3 A, 6.8 A]. When η is set as 0.15 and 0.11, the chaotic behavior can be stabilized into a period-1 orbit and a period-2 orbit, as shown in Figures 6.3b and 6.3c, which are in good agreement with the stable domains shown in Figure 6.2a. The corresponding boundaries of the period-1 trajectory are img [0.66 V, 1.9 V] and img [2.2 A, 5.8 A], and those of the period-2 are img [−0.5 V, 3.6 V] and img [1.4 A, 7.4 A].

Figure 6.3 Simulated phase portraits. (a) Chaotic orbit. (b) Period-1 orbit. (c) Period-2 orbit

img

img

img

6.1.4 Experimentation

According to Figure 6.1, an experimental DC drive system is prototyped. In principle, there are three main subsystems, namely a power electronic DC chopper, a motor-generator set, and an analog electronic controller. The DC chopper, consisting of a DC power supply img, a power MOSFET switch IRFI640G, a power diode BYW29E200, and an inductor L, functions to regulate the input power flowing into the drive. The motor-generator set includes a DC motor, a DC generator, a coupler, and an electronic load, where the mechanical load torque is electronically controlled by the current sink of the electronic load. The electronic controller involves simple hardware, namely an encoder M57962L, a F/V converter LM331, three op-amps (OA1, OA2, and OA3) LM833, a bucket-brigade delay (BBD) line MN3004 and its clock MN3101, a ramp-signal generator, a comparator (CM) LM311, and a MOSFET driver DS0026.

Based on the encoder and the F/V converter, the motor speed ω is converted into an analog signal img (with gain γ) which is then compared with the command speed signal img to produce the error signal img via OA1 with gain α. Hence, the speed feedback gain g equals αγ. According to the principle of delayed self-controlling feedback (Pyragas, 1992), img and its delayed version img are fed into OA2 with gain β to produce the perturbation signal img which is then compared with img to generate the desired control signal img via OA3 with gain unity. Finally, img is compared with the ramp signal img (with period T and upper and lower bound voltages img and img) via the CM to produce the PWM switching signal for driving the power MOSFET. The core of this controller is the BBD line and the associated clock. By tuning the clock frequency via its externally connected R1R2C network, the BBD line can allow for a time delay varying from 2.56 to 25.6 ms. In the controller, the time delay τ is set to the switching period T which in fact corresponds to the fundamental operation (the period-1 orbit).

Based on the same conditions for computer simulation, the measured chaotic, period-1, and period-2 phase portraits are shown in Figure 6.4. It can be found that the chaotic orbit has boundaries of img and img, the period-1 orbit has boundaries of img and img, and the period-2 orbit has boundaries of img and img. Comparing Figures 6.4 and 6.3, the measured results are in good agreement with the simulation results. Nevertheless, the boundaries of those phase portraits still have some discrepancies, and the stabilized orbits are slightly shaking, which is due to some inevitable imperfections in the DC drive system, such as the uneven contacts of the DC commutator, the torsional oscillation of the coupler, and the phase distortion of the BBD line.

Figure 6.4 Measured phase portraits. (a) Chaotic orbit. (b) Period-1 orbit. (c) Period-2 orbit

img

img

img

6.2 Stabilization of Chaos in AC Drive System

As discussed in Chapter 4, chaotic behavior can be observed in the permanent magnet synchronous motor (PMSM) drive system under some operating conditions. By defining

img

where L is the armature winding inductance, R is the armature winding resistance, B is the rotor viscous friction coefficient, J is the rotor inertia, img is the number of PM pole-pairs, img is the PM flux linkage, img is the d-axis input voltage, img is the q-axis input voltage, and img is the load torque, the dynamical equation of the PMSM drive can be transformed into a dimensionless model which is in the form of the well-known Lorenz system (Hemati, 1994):

(6.39) equation

where img, img, and img are the transformed versions of the d-axis armature current img, q-axis armature current img, and motor speed ω, respectively, and img is a free parameter.

This section presents an overview of four control methods to stabilize the chaotic motion in a PMSM drive system. Namely, the nonlinear feedback control, backstepping control, dynamic surface control, and sliding mode control are employed to stabilize the chaotic motor speed to the desired value.

6.2.1 Nonlinear Feedback Control

A nonlinear feedback control has been proposed to stabilize chaos in a PMSM drive system (Ren and Liu, 2006). This nonlinear feedback control uses img and img as the manipulated variables. The control law is governed by:

(6.40) equation

(6.41) equation

where img and img are the control objectives of img and img respectively, while img and img are the corresponding controller parameters. By substituting (6.40) and (6.41) into (6.39), the dynamical equation of a PMSM drive system with a nonlinear feedback control is rewritten as:

(6.42) equation

Since img and img are normally constant parameters, this yields img and img. Thus, the control law can be reduced as:

(6.43) equation

(6.44) equation

The desired speed reference img is a constant for the PMSM drive system. Therefore, in the controller design, img can be calculated according to the given value of img and img is set based on the requirement of flux. The selection of img and img determines the responding speed of the system. The manipulated variables img and img are externally accessible, while img, img, and img can be measured in real time. Hence, the controller can be physically realized.

In order to assess the performance of this nonlinear feedback control, the simulation is carried out with the parameters img = 0, img = 5.46, img = 20. Given img = 5, it results in img = 5 according to img, and img = 3.

The simulation is carried out with the initial values img = 20, img = 0.01, and img =− 5 as well as img. After startup, img and img are set to zero and remain unchanged. At the instant of t = 25 s, img and img are adjusted in accordance with the control law governed by (6.43) and (6.44). As shown in Figure 6.5, it can be found that the nonlinear feedback control can successfully stabilize both of the armature current components and the motor speed.

Figure 6.5 Stabilization performance of nonlinear feedback control. (a) Transformed d-axis armature current. (b) Transformed q-axis armature current. (c) Transformed motor speed

img

img

img

6.2.2 Backstepping Control

The backstepping control method has also been designed to stabilize the chaotic motion in a PMSM drive system (Harb, 2004). By setting img = 0 and img = 0, the dynamical equation (6.39) of the PMSM drive is expressed as:

(6.45) equation

The error signals of the corresponding system states are defined as:

(6.46) equation

where img, img, and img are the gains of error signals. Given img, and by the use of (6.45), the time derivative of (6.46) can be represented as:

(6.47) equation

which constitutes the control law for stabilization.

Based on these error signals, a positive Lyapunov function is constructed as:

(6.48) equation

By selecting the control parameters img, img and img, the system stability is guaranteed, namely img, provided that the control law is given by:

(6.49) equation

Based on the same system to assess the nonlinear feedback control, the stabilization performance of this backstepping control is testified. Also, the control law is applied at the instant of t = 25 s. As shown in Figure 6.6, it can be seen that the backstepping control can successfully stabilize the system.

Figure 6.6 Stabilization performance of backstepping control. (a) Transformed d-axis armature current. (b) Transformed q-axis armature current. (c) Transformed motor speed

img

img

img

6.2.3 Dynamic Surface Control

The dynamic surface control has also been used to stabilize chaos in the PMSM drive system. It takes the advantage over the backstepping control because it can avoid the problem of ‘explosion of terms’ caused by the repeated differentiation of virtual input (Wei et al., 2007).

By using the control variable img and setting img in (6.39), the corresponding dynamical equation becomes:

(6.50) equation

The control law of this dynamic surface control includes three steps. Namely, the virtual controllers are designed in the first step and the second step, and the overall control law is designed in the third step (Wei et al., 2007).

Firstly, the dynamic surface img is designed for img to track the reference value img, which is expressed as:

(6.51) equation

Hence, by differentiating (6.51) and using (6.50), the dynamics of img is given by:

(6.52) equation

Since img is a constant value and img, (6.52) becomes:

(6.53) equation

So, the first virtual controller is to stabilize img by choosing img as:

(6.54) equation

where img is the parameter of the first virtual controller. It should be noted that the dynamic surface control eliminates the need for model differentiation by passing img through a first-order filter with a positive time constant img:

(6.55) equation

where img serves as an estimation of img.

Secondly, by using img to supersede img, the second surface img is designed as:

(6.56) equation

Hence, by differentiating (6.56) and using (6.50), the dynamics of img is given by:

(6.57) equation

So, the second virtual controller is to stabilize img by choosing img as:

(6.58) equation

where img is the parameter of the second virtual controller. Consequently, img is deduced by passing img through a first-order filter with a positive time constant img:

(6.59) equation

Thirdly, the third surface img is designed as:

(6.60) equation

So, by differentiating (6.60) and using (6.50), the dynamics of img is given by:

(6.61) equation

In order to stabilize img, the final control law is designed as:

(6.62) equation

where img is the parameter of the final control law.

Based on the same system as mentioned previously, the stabilization performance of this dynamic surface control is assessed. When img is selected and the control is activated at the instant of t = 25 s, the responses are as shown in Figure 6.7. This confirms that dynamic surface control can successfully stabilize chaos in a PMSM drive system.

Figure 6.7 Stabilization performance of dynamic surface control. (a) Transformed d-axis armature current. (b) Transformed q-axis armature current. (c) Transformed motor speed

img

img

img

6.2.4 Sliding Mode Control

The sliding mode control has also been developed to stabilize the chaotic behavior in a Lorenz chaotic system, such as a PMSM drive system (Yau and Yan, 2004). By setting img and img in (6.39), the dynamical equation of a PMSM drive system can be expressed as:

(6.63) equation

where img is a nonlinear control variable that is a continuous nonlinear function with img and follows the relationship:

(6.64) equation

where img and img are nonzero positive constants. A disturbance term is defined to cancel the following nonlinear term:

(6.65) equation

So, the system dynamics in (6.63) can be expressed as:

(6.66) equation

For analysis, a new system state vector img is defined, where the transformation matrix is img. The trajectory error states are defined as img, img, and img, where the transformed regulation system states are img and img. Thus, the error state dynamical equations are given by:

(6.67) equation

where img. So, a sliding surface suitable for the application can be defined as:

(6.68) equation

where img is the design parameter. For a sliding mode operation, the necessary and sufficient conditions are given by:

(6.69) equation

(6.70) equation

Therefore, the sliding mode dynamics can be obtained as:

(6.71) equation

(6.72) equation

(6.73) equation

When the design parameter img, the stability of (6.71) is guaranteed, and img converges to zero. According to (6.69), img is also stable and converges to zero. It has been proved that img will converge to zero if img and img converge to zero (Yau and Yan, 2004). Therefore, (img) converges to the desired value (img) when img.

The sliding mode control law is designed as:

(6.74) equation

where

(6.75) equation

In this way, a reaching condition for the sliding mode img can be guaranteed (Yau and Yan, 2004).

In order to testify the performance of the sliding mode control, the simulation is carried out with the system parameters img, img, and img. The design parameter img is selected, while the nonlinear input is defined as:

(6.76) equation

Thus, the slope of the nonlinear sector img and img can be obtained, leading to the selection of img. Based on (6.67), (6.74), and (6.76), the simulated responses of img, img, and img under img are depicted in Figure 6.8 in which the control is activated at the instant of t = 25 s. This confirms that a sliding mode control can successfully stabilize chaos in a PMSM drive system.

Figure 6.8 Stabilization performance of sliding mode control. (a) Transformed d-axis armature current. (b) Transformed q-axis armature current. (c) Transformed motor speed

img

img

img

References

Boccaletti, S., Farini, A., and Arecchi, F.T. (1997) Adaptive strategies for recognition, control and synchronization of chaos. Chaos, Solitons and Fractals, 8, 1431–1448.

Casas, F. and Grebogi, C. (1997) Control of chaotic impacts. International Journal of Bifurcation and Chaos, 7, 951–955.

Harb, A.M. (2004) Nonlinear chaos control in a permanent magnet reluctance machine. Chaos, Solitons, and Fractals, 19, 1217–1224.

Hemati, N. (1994) Strange attractors in brushless DC motors. IEEE Transactions on Circuits and Systems - I: Fundamental Theory and Applications, 41, 40–45.

Hunt, E.R. (1991) Stabilizing high-period orbits in a chaotic system – the diode resonator. Physics Review Letters, 67, 1953–1955.

Hirasawa, K., Wang, X.F., Murata, J. et al. (2000) Universal learning network and its application to chaos control. Neural Networks, 13, 239–253.

Jackson, E.A. and Grosu, I. (1995) An OPCL control of complex dynamic systems. Physica D, 85, 1–9.

Khovanov, I.A., Luchinsky, D.G., Mannella, R., and McClintock, P.V.E. (2000) Fluctuations and the energy-optimal control of chaos. Physics Review Letters, 85, 2100–2103.

Kittel, A., Parisi, J., and Pyragas, K. (1995) Delayed feedback control of chaos by self-adapted delay time. Physics Letters A, 198, 433–436.

Liao, T.L. and Lin, S.H. (1999) Adaptive control and synchronization of Lorenz systems. Journal of Franklin Institute - Engineering and Applied Mathematics, 226, 925–937.

Ott, E., Grebogi, C., and Yorke, J.A. (1990) Controlling chaos. Physical Review Letters, 64, 1196–1199.

Poznyak, A.S., Yu, W., and Sanchez, E.N. (1999) Identification and control of unknown chaotic system via dynamic neural networks. IEEE Transactions on Circuits and Systems - I: Fundamental Theory and Applications, 46, 1491–1495.

Pyragas, K. (1992) Continuous control of chaos by self-controlling feedback. Physics Letters A, 170, 421–428.

Rajasekar, S., Murali, K., and Lakshmanan, M. (1997) Control of chaos by nonfeedback methods in a simple electronic circuit system and the FitzHugh-Nagumo equation. Chaos, Solitons and Fractals, 8, 1545–1558.

Ramesh, M. and Narayanan, S. (1999) Chaos control by nonfeedback methods in the presence of noise. Chaos, Solitons and Fractals, 10, 1473–1489.

Ren, H. and Liu, D. (2006) Nonlinear feedback control of chaos in permanent magnet synchronous motor. IEEE Transactions on Circuits and Systems - II: Express Briefs, 53, 45–50.

Tanaka, K., Ikeda, T., and Wang, H.O. (1998) A unified approach to controlling chaos via arm LMI-based fuzzy control system design. IEEE Transactions on Circuits and Systems - I: Fundamental Theory and Applications, 45, 1021–1040.

Tian, Y.P. (1999) Controlling chaos using invariant manifolds. International Journal of Bifurcation and Chaos, 72, 258–266.

Wei, D.Q., Luo, X.S., Wang, B.H., and Fang, J.Q. (2007) Robust adaptive dynamic surface control of chaos in permanent magnet synchronous motor. Physics Letters A, 363, 71–77.

Yau, H.T. and Yan, J.J. (2004) Design of sliding mode controller for Lorenz chaotic system with nonlinear input. Chaos, Solitons and Fractals, 19, 891–898.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset