You are on page 1of 6

2010 8th IEEE International Conference on Control and Automation Xiamen, China, June 9-11, 2010

ThCP4.12

Piezoelectric Self-Sensing Actuator for Vibration Suppression Based on Time-Sharing Method


First A. Bo Wang, Second B. Rongxiu Wang.

AbstractA new time-sharing self-sensing method using piezoelectric actuators for active vibration suppression is proposed and investigated. This self-sensing technique is based on the idea that one PZT can have two electrical properties, namely whether it acts as a sensor or as an actuator. The actuation characteristics of the time-sharing self-sensing piezoelectric actuators were discussed. The shift between a sensor channel and an actuation channel results in a chopped control force. By applying the method of independent modal space control, a new procedure for the control force optimization of this type actuator is introduced. A numerical vibration suppression simulation demonstrated that the time-sharing self-sensing method works well on a beam structure. An experiment was carried out to test the practicability and the results proved the validity.

I. INTRODUCTION

CTIVE vibration control seeks to damp unwanted structural vibrations by the application of an appropriate counter force. Piezoceramics are potential actuators for a wide range of applications in aerospace, automotive, civil structures, machine tools and big-medical systems to actively control vibration and noise, improve performance, and augment stability[1]. Typically, piezoceramics are used as actuators and polymer piezo films are used as sensing materials. It is also possible to use piezoceramics for both sensing and actuation. The motivation for a self-sensing actuator is that it is capable of truly collocated control. Collocated control is popular in active vibration control, and is a broad term that encompasses any control strategy where the control variable is sensed at the point of actuation. In practice, it is often difficult to sense and actuate at the same spot since this typically requires two separate devices to occupy the same space. However, a self-sensing actuator eliminates this problem. In addition, a self-sensing actuator is attractive from the standpoint of minimizing the amount of instrumentation required to control a structure. An actuator/sensor pair in a single package cuts instrumentation in half and reduces the number of system elements subject to failure for structural control applications such as large space structures where fault

Manuscript received August 25, 2009. F. A. Bo Wang is with the Chongqing University of Technology, Chongqing, 400050, China(phone: 023-68655596; fax: 023-68655596; e-mail: quantumfox@live.cn). S. B. Rongxiu Wang is with Chongqing T&B University, Chongqing, 400067, China(e-mail: wrx_07@hotmail.com).

tolerance and system redundancy is an issue. The first discussion of the idea took place in the 1950's when "motional feedback" was proposed as a means of damping mechanical resonances in loudspeakers[2]. Motional feedback is a concept that falls naturally out of modeling the transduction process as a two-part network. Essentially, it is built on the fact that the voltage across the electrical terminals of an electromechanical transducer is a "linear" combination of two voltages: one induced by the applied potential, the other induced as a result of motion in the mechanical system. If a means can be found to separate these two voltages, and measure the one due to mechanical motion, then you have a means of using an electromechanical transducer as both an actuator and sensor at the same time, i.e., a self-sensing actuator. Similarly, when a single patch (PZT) is used as a self-sensing actuator, it possesses many advantages over the use of two separate PZT elements as individual sensors and actuators. This concept of a self-sensing piezoelectric actuator was first developed and published by Dosch et al.[3]. Self-sensing piezoelectric actuator systems are lighter and less costly than non-collocated sensor/actuator systems. Moreover, the collocation of sensing and actuation allows the control signal to be applied at the point of measured response, thereby eliminating the capacitive coupling between the sensor and actuator elements[3]. Goh and Caughley presented results which demonstrate that structures controlled with collocated velocity feedback are unconditionally stable at all frequencies[4]. Subsequent to the work of Dosch et al., self-sensing piezoelectric actuators have been widely employed in vibration and control applications. Tzou and Hollkamp investigated the use of self-sensing orthogonal modal actuators to effectively control vibration in beam-like structures[5]. Frampton et al. and Dongi et al. investigated the feasibility of using a self-sensing actuator for active flutter suppression[6],[7]. Vallone successfully applied self-sensing actuators to control vibrations in large-scale space structures[8]. In addition, self-sensing actuation has been extended to active acoustic noise control[9],[10] and structural health monitoring applications[11],[12]. While most studies use monolithic piezoceramic material for its d31 coupling coefficients, Jones et al. applied the self-sensing actuation concept to PZT stack actuators acted as a micropositioner[13]. Sodano et al. investigated the feasibility of Macro-Fiber Composites used in self-sensing actuation for

978-1-4244-5196-8/10/$26.00 2010 IEEE

1403

ThCP4.12

vibration reduction in flexible structures, such as inflatable space devices[14]. The advantages gained through the use of a piezoelectric self-sensing actuator are well documented; however, some difficulties remain. The most popular ways to separate the control and sensing signals are bridge circuits by employing a single piece of piezoceramic material to simultaneously sense and actuate in a closed-loop system, yet maintaining the balance of the bridge circuit presents a great challenge. The high sensitivity of the bridge circuit to variations in the parameter values is a serious drawback. A bridge circuit can be imbalanced by a mere one per cent variation in piezoelectric capacitance, and this can destabilize active self-sensing control systems[15]. The capacitance of PZT is temperature sensitive, causing it to vary significantly as temperature changes. When used as an actuator, the voltage applied to the piezoelectric material tends to be much greater than that generated during sensing. Since both signals occur simultaneously in the material, it becomes very difficult to distinguish the sensor voltage from the mixed signals, as Tani et al. reported[16]. Furthermore, the bridge configuration increases power consumption and each PZT needs an independent bridge circuit, resulting in complex circuitry when several PZTs are applied. As a consequence of its potential commercial value, the development of robust or adaptive self-sensing actuators became a recurrent research topic, therefore, despite the drawbacks, various kinds of bridge circuit self-sensing approaches have been proposed and extensively investigated to improve the stability and the performance[17]-[24]. The authors proposed a new self-sensing approach using switching control, and verified its validity through the self-sensing vibration suppression experiments of a cantilever beam[25]. The self-sensing method was devised by the time-sharing method of a single piezoelectric actuator. This self-sensing technique is based on the idea that one PZT can have two electrical statuses concerning its electric current or voltage, namely whether it acts as a sensor or as an actuator, therefore, the two different functions of a PZT patch as a sensor or as an actuator can be selected by connecting the switch to its sensing electric network or to its actuation electric network. Fig.1 briefly shows how such a self-sensing actuator operates. It utilizes the piezoelectric effect and the piezoelectric converse effect by turns through the shift of a switch controlled by a time base signal from a computer or a controller, making one PZT act as a sensor in one time-span and act as an actuator in the next. When the switch is connected to the sensing signal channel the state of vibration can be estimated with the piezoelectric voltage or charge; whereas the switch is connected to the control signal channel an electric field is applied to drive the PZT actuator. In such cases one PZT patch can be used as both a sensor and an actuator alternately. Although this basic idea and procedure of implementation were described in details previously, its distinguishing features still need to be theoretically analyzed

to obtain optimization performance with the minimization of control spillover. Usually, the spillover problem can be dealt with filtering techniques and independent modal space control method[26]. The work presented in this paper is an attempt to discuss the control force characteristics of this time-sharing self-sensing actuator and set up a specific procedure for the optimization. To achieve the optimization force, the modal space control method is applied, and a suitable optimization criterion is formulated. In the end, numerical examples and an experiment are presented for a cantilever beam with self-sensing piezoelectric actuators to demonstrate the effectiveness and validity of the time-sharing self-sensing approach.
Sensing signal

Controller

Fig. 1. The operation of the self-sensing actuator.

II. THE ACTUATION CHARACTERISTICS OF THE TIME-SHARING SELF-SENSING ACTUATORS Consider a linear system described by

MZ + CZ + KZ = bf c (1) where M is the mass matrix of the system, C is the systems damping matrix, K the stiffness matrix, b the controlling influence matrix decided by the locations of the actuators, fc the control vector that are the forces generated by actuators or control voltages.
0.5

Continuous force

0.4 0.2 0 -0.2 -0.4


t(s)

Discontinuous force

PZT cell

Switch Control signal Switching control signal

Ground

0.5

1.5

-0.5

t(s)
0 0.5 1 1.5 2

Fig. 2. Smooth continuous force and chopped force. The black bars indicate that a PZT cell is acting as an actuator to provide control force during a short time interval, while the blanks indicate the same PZT cell acting as a sensor.

The control force generated by a piezoelectric actuator is often assumed to be continuous and smooth if its voltage varies continuously and smoothly. However, it is not the case when a time-sharing self-sensing piezoelectric actuator is applied due to the disconnection of the switch from control signal electric circuit for a short period of time, as shown in Fig.2. Thus, the control force provided by a time-sharing self-sensing actuator becomes chopped or intermittent, but keep constant before it is shifted to sensing line again. Those features demand a significantly different way to design its optimal control force. By Z = q , (1) can be expressed as follows

1404

ThCP4.12

q + Dq + diag (

2 i

)q = bf c = f

(2)

III. THE DETERMINATION OF CONTROL FORCE When piezoelectric PZTs are used as time-sharing self-sensing actuators, as discussed above, their control forces become chopped, keep constant during a short time period, as shown in more details in Fig.3. In order of simplification, consider the ith mode and its behavior in one time cycle T . Assume its state vectors are qi (t j ) , qi (t j ) at time tj. Divide the time interval from tj to

where is the system eigenvectors, i is the ith eigenvalue. If C can be decoupled the matrix D will be diagonal, then (2) stands for l separated one dimensional equations. qi + d ii qi + i2 qi = f i (i=1,2,3, l) (3) where dii is the element of D in ith row and ith column, l is the dimension of modal space. For the ideal continuous and smooth control force, the ith modal response at time t is expressed as t f ( ) qi = i e ni ( t ) sin pi (t )d (4)
0

tj+1 into two parts, denote them as 1 and 2 respectively, and T = 1 + 2 = tj+1-tj, see Fig.3. The structure under controlled
vibrates freely with damping during 1 and the piezoelectric patch acts as sensor, while it acts as actuator during 2 to provide a constant control force fi. Assign tj=0, then the equations of motion of the ith mode during tj+1-tj and after can be expressed as qi + d ii qi + i2 qi = 0 (0<t 1 ) (7)

where ni=dii/2, p = n . If N is a large number (4) can be approximately transformed into N f (t ) n ( t t ) qi = i j e sin pi (t t j )T (5)
2 i 2 i 2 i
i j

j =1

where N=t/ T , T =tj+1-tj= 1 + 2 . When the intermittent control force mentioned above is considered, and note that f i (t j ) = 0 between tj and tj+ 1 by Fig.3, the response of the ith mode approximately becomes N f (t ) n ( t t ) qi = i j e sin pi (t t j ) 2 T (6) i 1 + 2 j =1 Compared with (5), (6) gets a fraction of the value. This result clearly indicates that the system driven by time-sharing self-sensing actuators responses differently to the same excitation voltage. The optimized control forces designed in common cases may not be optimal for time-sharing self-sensing piezoelectric actuators.
i j

qi + d ii qi + i2 qi = f i
2 i

( 1 <t 1 + 2 )

(8) (9) and

qi + d ii qi + qi = 0 (t> 1 + 2 )

where the initial conditions are qi (0) = qi (t j )

qi (0) = qi (t j ) . The solutions to (7), (8) and (9) lead to qi (t ) = Ai e n t sin( pi t + i ) where Ai2 = ( Ai0en ( + ) )2 + Bi fi 2 + [Ci qi (0) + Di qi (0)] fi
i i 1 2

(10) (11)

Bi =
Ci =
Di = 2 p i
2 i

1 (1 + e 2 n 2e n cos i 2 ) pi2i2
i 2 i 2

2 i

e ni (1 + 2 ) cos i ( 1 + 2 )
i 1 2

i2

e ni (1 + 2 2 ) cos i 1

fi fj tj

n 2 n ( + ) [sini ( 1 + 2 ) + i cosi ( 1 + 2 )] e i pi2i e n ( +2 ) (sini 1 +


i 1 2

fj+1 tj+1

fj+2 tj+2 t(s)

ni

cosi 1 )

Fig. 3. Time sharing process of a self-sensing piezoelectric actuator. During 1 periods, no electric field is applied on the PZT cell, it acts as a sensor while in

tg ( i ) =

A e ni 2 sin(i 2 + ) + A e ni 2 cos(i 2 + ) +
2

fi pi2 fi pi2

ni

(12)

periods it is driven by a constant voltage, acting as an actuator.

A = ( Ai 0 e ni1 ) 2 +

Since control signals lag behind the states of vibrating structures due to signal procession the time delay problem often exists in active feedback vibration control systems. This problem is usually ignored if it does not obviously cause instability and spillover. However, the suppression performance might be greatly degraded by its intrinsic time-delay in time-sharing self-sensing actuation systems. Thus it is another factor that should be taken into consideration. In order to overcome these drawbacks, we introduce a simple way to determine the control force.

fi 2f i cos i 1e ni1 qi (0) pi2i2 i2

ni 1 2 fi n (sin i 1 + i cos i 1 )e qi (0) 2 pi i i

Ai 0e ni 1 sin( i 1 + i 0 ) tg ( ) = Ai 0e ni 1 cos( i 1 + i 0 ) ni

fi pi2 fi pi2

Ai0 and i 0 are initial amplitude and angle at time ti, they are

1405

ThCP4.12

Ai 0 = q(t j ) +
2

(q(t j ) + ni q(t j ))2 pi2 .

and

tg (i 0 ) =

q(t j ) pi q(t j ) + ni q(t j )

vibration mode f i * depends on the structural vibration state as well as the three parameters Ci, Bi, and Di which in turns are dependent on the values of 1 and 2 . To keep the control force with the pace of the variation state, the smaller the values of 1 and 2 , the better. Larger 1 and 2 indicates more time-delay. However, the values of 1 and 2 are usually limited by electrical network properties of a real control system and the needs of signal process, say A/D conversion, sampling, or slew rate of the amplifier. Besides, smaller 1 and 2 demands higher quality of hardware. Nevertheless the determination of 1 and 2 is arbitrary to some extent. As we can find out later in the numerical example that the variations of 1 and 2 cause little change to the system performance. It is easy to see that Ci, Bi, Di are 2 all cyclic functions of 1 and 2 , thus we have 1 + 2 <

According to (11), it is possible to effectively suppress the ith modal vibration amplitude by properly selecting the control force fi and the parameters Ci, Bi, and Di. When 2 or fi is zero, (11) and (12) represent the vibration amplitude and phase angle without control, at such case Bi=Ci=Di=0. If 1 =0, this condition means that the PZT cell acts as an actuator all the time, that is the case of conventional piezoelectric actuators. To obtain the suitable control force, we take the derivative of Ai with respect to fi. dAi2 = 2 Bi f i + Ci qi (0) + Di qi (0) df i d 2 Ai2 = 2 Bi df i 2 dAi2 = 0 , which leads to df i (13) (14)
2 2

Since Bi is always nonnegative the minimum of Ai can be obtained by assigning Ai2 = ( Ai 0e n ( + ) ) 2


i 1 2

which can be easily satisfied in most cases. We also note, with the help of (12), that phase angle of the ith mode shifts because of the exertion of the force fi. We did not discuss this phenomenon due to its less importance, but it will be clearly shown in our later numerical simulation. V. NUMERICAL SIMULATIONS AND EXPERIMENT To investigate how the self-sensing method works, numerical simulations were performed. Take a cantilever beam with PZTs the numeric example, as Fig. 4 shows.
PZT1 PZT2 PZT3 PZT4

[Ci qi (0) + Di qi (0)]2 (15) 4 Bi The optimal control force of fi is C D f i * = i qi (0) i qi (0) (16) 2 Bi 2 Bi Note that (16) takes a form similar to that of LQR. In (15), the first term represents the decreased amplitude of vibration caused by damping within each cycle T = 1 + 2 and the second term represents the influence of control force fi. The control force for the ith mode at period 2 can be calculated provided the state of vibrating structure at period 1 , the modal displacement and velocity, and the parameters Ci, Bi, and Di are known. This approach theoretically eliminates the probable instability resulted from time-delay. The independent modal space method minimizes the control spillovers if the eigenvalues and eigenvectors are known accurately and the vibration modes of the structure are not dense[26]. In practice, not all of the vibration modes of the structure are necessarily taken into account, and we can select vibration modes to be suppressed. IV. SOME CONSIDERATION ABOUT THE TIME CYCLE In the previous section we discussed the actuation of a time-sharing self-sensing actuator, and tried to set up a criterion for the control force optimization of the ith mode. According to (16), the required control force for the ith

Fig. 4. Configuration of a beam with PZTs attachment at 0mm, 1000mm, 1500mm, and 2250mm according to the Maximal Modal Force Rule. MMFR indicates that the maximal modal forces for the lowest four modes appear at x=0, x=l/2, x=l/3 and x=3l/4 respectively, and the length of the beam l=3000mm.

The properties of the beam and of the piezoelectric patches are as follows. For the beam, size 3000mm 50mm10mm, Youngs modulus 4.51012Pa and the density 5500kg/m3. For the PZT, Size 2.5cm2cm1mm, density 7500kg/m3, Youngs modulus 6.31010Pa, piezoelectric coefficient 1.210-10V/m, rated voltage load 125V. PZT patches are all bonded on the surface of host beam. The lowest four modes are used to approximate the beams deflection, and are selected to be controlled. Four self-sensing piezoelectric actuators are used accordingly. The locations of the PZT patches are determined by Maximal Modal Force Rule (MMFR)[27], and they locate at x=0, x=l/2, x=l/3 and x=3l/4 respectively, where l is the length of the beam, that is l=3000mm. For the lowest three modes, the highest angular 3 is 89.0(rad/s), so we frequency choose 1 = 0.01s , 2 = 0.025s . The fourth mode has a much

1406

ThCP4.12

higher

angular

frequency,

4 = 176.3

(rad/s),

the changes more obviously in Fig.5 and Fig.6.

so 1 = 0.005s , 2 = 0.01s are set for it. The vibration suppression simulation results are shown in Fig.5-Fig.8.

Fig. 5. Displacement and control voltage of the first mode( =5.1rad/s)

Fig. 8. Displacement and control voltage of the fourth mode( =176 rad/s, 1=0.005s, 2=0.01s)

Fig. 9 shows the time histories of the second mode vibration suppression under two different settings of 1 and 2 . Although the two curves are different the effective vibration suppression are similar. It demonstrated that the selection of the values of 1 and 2 is flexible provided we know the limitation of system hardware.

Fig. 6. Displacement and control voltage of the second mode( =32.0 rad/s)

Fig. 9. Control results at different values of 1 and 2 for the second mode with =32.0(rad/s). The two curves show a little bit different vibrating phases, but the overall shape look alike.

Fig. 7. Displacement and control voltage of the third mode( =89.0 rad/s)

Fig. 5 to Fig. 8 show time histories of the four lowest mode vibration suppressions and their applied control voltages with the time-sharing self-sensing method. During the intervals of the sensing, the voltages are zero and the piezoelectric actuators sense the state of structural vibration; while during the intervals of the actuation, the voltages are constant, as the self-sensing method specified in previous sections. By switching between the sensing channel and the actuation channel, a PZT patch operates as a sensor and an actuator alternately. Through this numerical simulation, it was confirmed that the time-sharing self-sensing approach worked well to suppress the vibration of a structure. The vibrating phases of the controlled structure are continuously changed by the application of chopped control forces. The variations are relatively larger for lower modes, as we can see

To test the feasibility of the new time-sharing self-sensing vibration suppression scheme, an experimental system was set up, similar to what we carried out in our previous work[25]. We used the same cantilever beam structure and PZT patches discussed in numerical simulation shown in Fig.4. The PZTs were customized by Sunnytec as described before. A commercially available MOSFET operation amplifier PA78 provided by Apex Microtechnology was used to drive the piezoelectric actuators. PA78 has a high slew rate, up to 350V/us, with an output of 175V and above 200 KHz operation speed. The high speed switch was the model ESC-RLY03-ISO produced by Leadertech which can work at 200V at 0.05ms to 0.25ms. The Advantech PCI-1710HG multifunctional A/D and D/A convertor was used, and its conversion time for A/D is 8us with a gain in the range of 0.5-2000, and the time for D/A conversion is 25us with a output voltage in 10V. It is programmable and can be plugged into computer. In this time, only the lowest two modes are selected to be controlled. The experimental result of the vibration suppression of the beam is shown in Fig.10.

1407

ThCP4.12

2
Displacement(cm)

Uncontrolled Controlled

1 0 -1 -2 0 0.5 1 1.5 2 2.5 t(s) 3

Fig. 10. The experimental result for the two lowest modes( =5.1 and 32.0 rad/s; 1=0.01s, 2=0.02s) vibration suppression compared with the uncontrolled vibration curve. Both signals were picked up at x=l/3.

VI. CONCLUSIONS The properties of piezoelectric materials, the piezoelectric effect and the piezoelectric converse effect can be utilized concurrently by time-sharing method to configure self-sensing actuators. This paper has presented a study in the actuation characteristics of such time-sharing self-sensing piezoelectric actuators. The direct applications of optimal control forces calculated for usual PZT actuators result in different responses to such time-sharing self-sensing actuators. It was shown that the control forces applied by such actuators are discontinuous or chopped, a new method to calculate the control force was proposed under modal space. The numerical and experimental results show the validity of such an approach. REFERENCES
[1] I. Chopra, Review of current status of smart structures and integrated systems. SPIE smart Structures and Integrated Systems, 1996, 2717:20-62 [2] Egbert De Boer, Theory of motional feedback. IRE Transactions on Audio, pp 15-21, January-February, 1961. [3] J. J. Dosch, D. J. Inman, and E. Garcia, A self-sensing piezoelectric actuator for collocated control. Journal of Intelligent Material Systems and Structures, Vol.3, pp.166-185, 1992. [4] C. J. Goh and T. K. Caughey, The stability problem caused by finite actuator dynamics in the control of large space structures. International Journal of Control, Vol.41,No.3,pp.787-802,1985. [5] H. S. Tzou and J. J. Hollkman, Collocated independent modal control with self-sensing orthogonal piezoelectric actuators(theory and experiment). Smart Materials and Structures, Vol.3, pp.147-156, 1994. [6] K. D. Frampton, R. L. Clark, and E. H. Dowell, Active control of panel flutter with linearized potential flow aerodynamics. AIAA Paper No.95-1079, Proceeding of the 36th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, April 10-13th, New Orleans, LA, pp.2273-2280, 1995. [7] F. Dongi, D. Dinkler, and B. Kroplin, Active panel flutter suppression using self-sensing piezoactuators. AIAA Paper No.95-1078, Proceeding of the 36th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, April 10-13th, New Orleans, LA, pp.2264-2272, 1995. [8] P.Vallone, High-performance piezo-based self-sensor for structural vibration control. Proceedings of 2nd SPIE Smart Structures and Materials conference, Newport Beach, CA, Vol.2443, pp.643-655, 1995. [9] B. Ko and B. H. Tongue, Acoustic control using a self-sensing actuator. Journal of Sound and vibration. Vol.187, pp.145-165, 1995. [10] D. J. Leo and D. Limpert, Self-sensing technique for active acoustic attenuation. Journal of Sound and Vibration. Vol.235, No.5, pp.863-873, 2000.

[11] C. Pardo de Vera and J. A.Guemes, Imbedded self-sensing piezoelectric for damage detection. Journal of Intelligent Material Systems and Structures. Vol.9, No.11, 876-882,1999. [12] J. S. Vipperman, Simultaneous qualitative health monitoring and adaptive piezoelectric sensoriactuation. AIAA Journal, Vol.39, No.9, 1822-1825,2001. [13] L. Jones, E. Garcia, and H. Waites, Self-sensing control as applied to a stacked PZT actuator used as a micropositioner. Smart Materials and Structures. Vol.3, pp.147-156,1994. [14] H. A. Sodano, G. Park, and D. J. Inman, An investigation into the performance of macro-fiber composites for sensing and structuaral vibration applications. Mechanical Systems and Signal Processing, Vol.18, No.3, pp.683-697. [15] E. H. Anderson and N. W. Hagood, Simultaneous piezoelectric sensing/actuation analysis and application to controlled structure. J.Sound Vib, Vol.174,No.5 619-639,1994 [16] J. Tani, G. Cheng and J. Qiu, Effectiveness and limits of self-sensing piezoelectric actuators. Proceedings of 1st Structural Health Monitoring Workshop, Stanford, CA, 503-514, 1997 [17] Z. A. Chaudhry, T. Joseph, F.P. Sun and C. A. Rogers, Local-area health monitoring of aircraft via piezoelectric actuator/sensor patches. SPIE 1995 VOL.2443, pp.268-276 [18] J. M. Yellin and I. Y. Shen, A self-sensing active constrained layer damping treatment for a Euler-Bernoulli beam. J Smart Mater Struct, 1996, 5:628-637 [19] Dong Weijie, Sunbaoyuan, Cui Yuguo, and Yang Zhixin. The vibration control of a cantilever beam based on self-sensing PZT actuator, Bulletin of Da Liang University, 2000,41(1) 77 80 [20] E. S. Garnett.Jr., R. H. Jeffrey, D. M. David, G. Park, and H. Sohn, Improved piezoelectric self-sensing actuation. Journal of Intelligent Material Systems and Structures, 15(12):941-953,2004 [21] Kanjuro Makihara, Junjiro Onoda, and Kenji Minesugi. Novel approach to self-sensing actuation for semi-active vibration vibration suppression. AIAA Journal, Vol.44, No.7, 1445-1453,2006 [22] R. L. Spangler and S. R.Hall, Broadband active structural damping using positive real compensation and piezoelectric simultaneous sensing and actuation. Smart Materials and Structures, Vol.3, 448-458, 1994 [23] D. G. Cole and R. L. Clark, Adaptive compensation of piezoelectric sensoriactuators. Journal of Intelligent Materials Systems and Structures. Vol.5, 665-672, 1994 [24] L. D. Jones and E. Garcia, Novel approach to self-sensing actuation. Proceedings of SPIE, Vol.3041,305-314, 1997 [25] Wang Bo and Yin Xuegang, A new type of self-sensing actuator and its application. Proceedings of the 5th International Conference on Vibration Engineering, Nanjing, China, 2002, 743-749 [26] Gu zhongquan, Ma kougen, and Chenweidong, The active vibration control, PeKing:National Military Industrial publisher, 1997 [27] Li Bin, Li Yugang, Yin Xuegang, and Huang Shanglian, Maximal modal force rule for optimal placement of point piezoelectric actuators for plates. Journal of Intelligent Material Systems and Structures, Vol.11,No.(5), 512-515,2000

1408

You might also like