You are on page 1of 9

This article has been accepted for inclusion in a future issue of this journal.

Content is final as presented, with the exception of pagination.


IEEE TRANSACTIONS ON ENERGY CONVERSION 1

Damping of Torsional Vibrations in a Variable-Speed Wind Turbine


John Licari, Student Member, IEEE, Carlos E. Ugalde-Loo, Member, IEEE, Janaka B. Ekanayake, Senior Member, IEEE, and Nicholas Jenkins, Fellow, IEEE

AbstractTorsional dampers are employed in wind turbines to damp vibrations in the drive-train. The conventional design is based on band-pass lters (BPF); however, its effectiveness can be compromised due to parametric uncertainty. To restore the performance of the damper, it is a common practice to re-tune it during the commissioning of wind turbines. To overcome this shortcoming, a model-based torsional damper was designed and its performance compared to the conventional approach when subjected to model uncertainty. A stability analysis was conducted and simulations were performed in Simulink. A real-time hardware-in-theloop experiment was carried out, with experimental and simulation results showing good agreement. The proposed model-based torsional vibration damper showed a superior performance over the conventional BPF-based approach. Results also showed that the model-based damper can eliminate the need for retuning procedures associated with BPF-based designs. Index TermsBand-pass lter (BPF), Kalman lter (KF), model based, real-time implementation, stability, torsional vibration damper, uncertainty, wind turbine.

I. INTRODUCTION

S the size of wind turbines increases, manufacturers face more challenges including achieving better cost effectiveness. This is currently being addressed through reduction of component masses leading to turbines becoming less tolerant of fatigue loads [1]. This study is concerned with the mitigation of fatigue loads in the drive-train of variable-speed wind turbines. Turbulent winds and gusts can excite modes that can lead to torsional vibrations in the drive-train, which in turn produce large stresses on components. Ultimately, this may reduce the lifetime of components, such as the gearbox [2], [3]. Moreover, when torsional vibrations are present in the drive-train, they will be converted to electrical power oscillations. This is highly undesirable for the operation of the power system because such oscillations can interact with the power system modes and resonance may result [4], [5]. Torsional vibrations can be suppressed by including a torsional vibration damper in the wind turbine drive-train. This helps to reduce mechanical stresses in the drive-train and avoid

Manuscript received June 6, 2012; revised August 15, 2012; accepted October 5, 2012. This work was supported in part by Nordic Windpower and the Low Carbon Research Institute. Paper no. TEC-00285-2012. The authors are with Cardiff University, Cardiff, CF24 3AA, U.K. (e-mail: LicariJ@cardiff.ac.uk; Ugalde-LooC@cardiff.ac.uk; EkanayakeJ@cardiff. ac.uk; JenkinsN6@cardiff.ac.uk). Digital Object Identier 10.1109/TEC.2012.2224868

potential resonances with the power system. It is important to damp these vibrations at all wind speeds; however, they are most signicant at high wind speeds during which the generator torque is held constant and the pitch controller is active [6]. Although the torsional vibration damper presented in this study works well for both low and high wind speeds, results of a wind turbine at above-rated wind speed are shown. This study was performed on a variable-speed wind turbine equipped with a full-rated converter (FRC); however, these results apply equally to other types of variable-speed technologies, such as those based on doubly-fed induction generators (DFIG) [7]. In such systems, there is a higher risk of exciting torsional vibrations from grid disturbances because the stator is connected directly to the electrical grid [8]. When insufcient vibration damping is present in a DFIG-based wind turbine, the vibrations can exhibit an exponentially increasing magnitude if not controlled [9]. Damping of torsional oscillations in the drive-train can be accomplished by either the pitch or the generator torque controller. However, it has been reported that the torque controller is more effective for this task [10] and therefore it was employed in this paper. A torsional vibration damper was included in the torque loop. This generates a damping torque which is added to the generator torque reference to provide an auxiliary damping to the system [6], [11][15]. A well-known torsional vibration damper design approach based on band-pass lters (BPFs) has been successfully adopted in many wind turbines [12][14]. The stability and performance of a system equipped with a BPF-based damper can be compromised if subjected to model uncertainties (e.g., discrepancies in the wind turbine model used for a control design and the actual wind turbine). In such a case, retuning of the damper is necessary to ensure a good damping performance [16]. An alternative approach that has been suggested for rolling mill applications involves a model-based damper. Here, damping is achieved through pole placement using state feedback with state estimation [17]. This damper design approach can be implemented on all variable-speed wind turbines when torque control is performed on the generator. A BPF-based torsional vibration damper was rst designed and its performance tested when subjected to model uncertainties. A model-based torsional damper was then designed in order to overcome the problem of retuning associated with the BPF approach. A stability analysis in the frequency domain was performed to assess the effect of the two different dampers on the wind turbine drive-train. Simulations to compare the performance of the vibration dampers were performed for a wind

0885-8969/$31.00 2012 IEEE

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
2 IEEE TRANSACTIONS ON ENERGY CONVERSION

equations referred to the low-speed shaft (LSS) are J1 d d rot = aero K1 (1 2 ) D1 (1 2 ) dt dt d 3 J2 2 = K1 (2 1 ) K2 2 + dt N D1 J3


Fig. 1. Block diagram of the wind turbine model implemented in Simulink.

(2)

d d (2 1 ) D2 dt dt 3 2 N D2

2 d dt

3 N

(3)

d gen = m K2 dt

3 2 N (4)

step and turbulent wind. To validate the simulation results, a hardware-in-the-loop experiment was carried out on a wind turbine test rig. II. WIND TURBINE MODEL A generic 2 MW variable-speed wind turbine based on a permanent-magnet synchronous generator (PMSG) with a threestage gearbox and an FRC was implemented in Simulink. The block diagram of the system is shown in Fig. 1. All parameters of the wind turbine model are given in Appendix A. A. Aerodynamic Model The aerodynamic model used is based on a look-up table which considers the power coefcient Cp as a function of the pitch angle and the tip-speed ratio . The aerodynamic torque is then calculated by [6] aero =
3 0.5AVwind CP (, ) rot

where N is the gearbox ratio. The damping coefcient was neglected in the model to have a worst case scenario. For this system, the mode frequencies of the blade in-plane and the drive-train are 2.54 and 3.7 Hz, respectively [18]. C. Electrical Model The electrical system consists of a back-to-back voltage source converter (VSC) and a PMSG. The VSC was modeled with ideal switches. The PMSG model (in the rotor reference frame) is described by [21] 1 Rs Lq d id = d id + npp gen iq dt Ld Ld Ld d 1 Rs Ld m npp gen iq = q iq npp gen id dt Lq Lq Lq Lq m = 3 npp [m iq + (Ld Lq ) id iq ] 2 (5) (6) (7)

(1)

where is the air density, A is the swept area by the rotor, Vwind is the wind speed, and rot is the speed of the rotor. The function of the aerodynamic model was mainly to initiate torsional vibrations in the mechanical model; thus, this simplied model was adequate [18]. B. Mechanical Model Torsional vibrations can be initiated directly by exciting the drive-train natural frequency mode or indirectly by exciting the blade in-plane symmetrical mode [12], [13]. Hence, a threemass model which considers both modes has been employed [18]. To simplify the rotor dynamics, the in-plane dynamics of the blades were represented as a torsional system [19], [20]. The three masses correspond to the inertia of the effective exible part of the blades J1 , the inertia representing the hub and the rigid part of the blades J2 , and the inertia of the generator J3 . The three inertias are separated by the effective blades stiffness K1 and the resultant stiffness of the low and high-speed shafts K2 , respectively [18]. The inputs to the mechanical model are the reaction torque of the generator m and the aerodynamic torque aero . The dynamic

where Ld and Lq are the d- and q-axis self-inductances, respectively, Rs is the stator winding resistance, id , vd and iq , vq are the d- and q-axis currents and voltages, respectively, m is the ux induced by the permanent magnets of the rotor into the stator windings, npp is the number of pole pairs, and gen is the generator mechanical speed. D. Control System The basic control objectives are to optimize the power production for below-rated wind speed and to limit the aerodynamic power for above-rated wind speed. For below-rated wind speed, the generator speed was varied by controlling the generator torque to follow an optimal torque versus speed curve derived from a generic 2 MW wind turbine model in Bladed. A vector control scheme was used to control the generator torque. For above-rated wind speed, the generator torque was held constant at the rated torque and the rotor speed was controlled by limiting the aerodynamic torque by varying the pitch angle. A pitch controller with gain scheduling was designed for this purpose [6]. It consisted of a PI controller with anti-windup, an actuator model represented by a rst-order lag, and a pitch rate and pitch angle limiters as shown in Fig. 2. The pitch controller parameters are given in Appendix B.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
LICARI et al.: DAMPING OF TORSIONAL VIBRATIONS IN A VARIABLE-SPEED WIND TURBINE 3

Fig. 2.

Pitch controller model.

Fig. 4.

Model-based torsional vibration damper.

frequency. The parameters of both the BPFs and the notch lter are included in Appendix B.
Fig. 3. BPF-based torsional vibration damper.

III. TORSIONAL VIBRATIONS DAMPING The basic controller outlined in Section II does not offer any mitigation to fatigue loads in the drive-train. A common industrial practice is to include a torsional vibration damper into the system [6]. Two different approaches have been used to design the torsional vibration damper: the rst one uses a conventional BPF and the other employs a model-based approach. A. BPF-Based Torsional Vibration Damper The structure of the damper used here consists of a notch lter cascaded with two BPFs and is shown in Fig. 3. A suitably tuned 2nd-order BPF was employed to extract the vibration frequency from the generator speed [12][14]. This information was used to add a small ripple at the vibration frequencies to the generator torque demand in order to damp torsional vibrations. The nominal mechanical frequencies associated with (2)(4) are relatively far apart (2.54 and 3.7 Hz); hence, two BPFs tuned at these frequencies were used [18]. The transfer function of each BPF is given by GBPF (s) = KBPF s2 2n s 2 + 2n s + n (8)

B. Model-Based Torsional Vibration Damper This alternative approach is shown in Fig. 4. The damping effect was obtained by carefully selecting the closed-loop location of the drive-train and blade in-plane resonant poles. This was achieved through pole placement (based on state feedback) [22]. In practice, state feedback is not realizable on its own; thus, an observer was used to estimate the state variables. In order to achieve good state estimation despite having uncertainty in the model and measurement noise, a discrete-time Kalman lter (KF) was used. A KF is a linear lter described by a set of mathematical equations [23]. These equations together with the three-mass model were included in the damper structure. The internal structure of the damper and the equations describing the KF are included in Appendix C. The KF minimizes the variance of the estimation error [23], [24]. Its successful implementation is ensured by having the three-mass model represented in the KF as accurately as possible. The closer the model within the KF structure is to the real plant, the better is the lter performance and thus the effectiveness of the damper. The damper performance was analyzed and tested for cases when the model included in the KF structure had the same dominant modes as the real plant and when there was a 10% of uncertainty. All input variables were available to the damper except the aerodynamic torque aero which had to be estimated. In the literature, this has been addressed by either augmenting the mechanical model with a state representing the unknown input [25], [26] or by solving the differential equation given by (2) using the derivative of the generator speed [27]. The derivative method was adopted in this paper. Although through this approach the aerodynamic torque estimate is obtained in a simple way, it is very sensitive to measurement noise. In order to mitigate the problem of noise associated with derivatives, a low-pass lter (LPF) was used. A compromise on the estimation delay introduced by the LPF and the noise rejected by the lter was reached [28]. All parameters of the model-based damper are included in Appendix B. The complete design procedure and the model-based damper inner control loop are given in Appendix C.

where is the damping ratio, n is the undamped natural frequency, and KBPF is the dc gain. Ideally, the BPFs must have a narrow peak (low damping ratio ) to limit their response at multiples of the blade passing frequency such as 3P or 6P and consequently avoid unwanted low-frequency variations in the torque and power in the system [6]. Since the wind turbine employed here had a 6P frequency of 1.8 Hz at rated wind speed (which is close to a resonant mode frequency, 2.54 Hz), a notch lter of the form GNF (s) =
2 s2 + 21 n s + n 2 + 2 s + 2 s 2 n n

(9)

was cascaded with the BPFs to restrict their response at multiples of the blade passing frequency. 1 and 2 are damping ratios which set the depth of the notch lter and n is the notch

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
4 IEEE TRANSACTIONS ON ENERGY CONVERSION

TABLE I CASES FOR DIFFERENT UNCERTAINTY LEVELS IN THE MODE FREQUENCIES

IV. STABILITY ANALYSIS OF THE TWO VIBRATION DAMPERS A feedback control system that can remain stable under a specied range of process perturbations is said to possess robust stability [29]. According to [30], practical experience suggests that a gain margin (GM) of 10 dB and a phase margin (PM) of roughly 60 are usually adequate stability margins. The Nyquist stability criterion was employed to evaluate system stability using the two vibration dampers. This criterion states that for a stable open-loop plant (such as the one representing the torque loop), closed-loop stability is maintained as long as there are no encirclements of the critical point (1, j0) [22]. Frequency domain methods, such as Nyquist analysis and Bode plots, are standard tools for analyzing linear time-invariant systems. It is well known that a wind turbine is a nonlinear system mainly due to the nonlinear aerodynamics [31]. For the stability analysis, only the torque loop shown in Fig. 1, which excludes the nonlinear aerodynamics, was considered. By further neglecting nonlinear terms in the three-mass model such as friction, the system can be assumed linear. Moreover, an additional simplication was performed by assuming that the electrical dynamics are at steady state when addressing the mechanical dynamics (as electrical dynamics of the system are much faster than the mechanical dynamics). Therefore, the electrical subsystem (generator and converter) was simplied and modeled as a rstorder lag, with time constant T [32], [33]. The simplied torque loops that were considered for the stability analysis are shown in Figs. 3 and 4. The stability analysis was carried out to identify regions of operation for which the system can potentially be unstable in the presence of model uncertainty. Uncertainty in the threemass model was included through variations of stiffness and inertia (Ji and Ki ) but not in the damping coefcients Di since they have no effect on the mode frequencies. Deviations in the mode frequencies for different values of Di (e.g., 0, 1 105 Nms/rad) were achieved through combinations of Ji and Ki . It was observed that realistic combinations of such parameter values had a negligible effect on system stability. System stability was assessed for different levels of uncertainties in the mode frequencies of the mechanical model. An uncertainty bound of 0.25 Hz (10% of F1 ) was introduced and nine cases were derived. These are tabulated in Table I. Case 5 represents a mechanical model without uncertainty which has
Fig. 5. Nyquist plots for the system with the BPF-based vibration damper.

Fig. 6.

Nyquist plots for the system with the model-based vibration damper.

nominal mode frequencies of 2.54 and 3.7 Hz. The Nyquist plots for the system with BPF-based and the model-based torsional vibration dampers for all cases in Table I are illustrated in Figs. 5 and 6. Fig. 5 shows that the phase and gain margins (PM = 34 , GM = 8 dB) of the mechanical model without uncertainty (Case 5) for the system with the BPF-based damper are below the recommended values [30]. Moreover, as the lower mode frequency uncertainty was increased negatively (Cases 13), the stability margins decreased drastically. The worst condition was in Case 1, where the system approached the critical point (1, j0), and as a consequence the stability margins deteriorated considerably (PM = 2.88 , GM = 1.09 dB). This result indicates that the Nyquist plot would encircle the critical point with a slight increase in either the damper gain or the level of uncertainty in the mechanical model, leading the system to instability. Although the system is stable for the given uncertainty bound (10%), its performance with the BPF-based damper is expected to deteriorate considerably when subjected to model uncertainty due to a large drop in the stability margins.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
LICARI et al.: DAMPING OF TORSIONAL VIBRATIONS IN A VARIABLE-SPEED WIND TURBINE 5

Fig. 7.

Closed-loop Bode plot for the system with the BPF-based damper.

Fig. 9.

Experimental platform.

Fig. 8.

Closed-loop Bode plot for the system with the model-based damper.

Fig. 6 shows that the Nyquist plot for Case 5 for the system with the model-based torsional damper features superior phase and gain margins (PM = 76.1 , GM ) over the BPF-based system. An increase in the uncertainty level produced a small change in the stability margins. The worst case was again Case 1; however, stability margins were still adequate (PM = 65 , GM ). Based on these attributes, the system with the modelbased damper is expected to perform better than BPF-based torsional damper. Besides that, the performance of the system with and without model uncertainty is not expected to change signicantly. A frequency domain analysis of the closed-loop system was performed using Bode plots for Cases 1 and 5. The closed-loop responses of the system with the different torsional vibration dampers are shown in Figs. 7 and 8. It can be seen that both dampers suppressed the resonant modes successfully with the nominal three-mass model (Case 5). However, for the case when frequency uncertainty was introduced (Case 1), the system with the BPF-based damper managed to damp the drive-train mode (3.95 Hz) but failed to damp the blade-in-plane mode (2.29 Hz). This is illustrated in Fig. 7. In the case of the system with the model-based damper, both frequency modes were damped successfully as shown in Fig. 8. V. EXPERIMENTAL PLATFORM The experimental platform used for the real-time system implementation is shown in Fig. 9. It consists of two components: a mechanically coupled motor-generator set and a back-to-back VSC. The control system was divided into a low and a highlevel controller. The low-level controller was implemented in

Fig. 10.

Experiment design block diagram.

a TMS320F2808 digital signal processor (DSP) and included the vector control schemes for the generator and grid-side converters. The high-level controller, implemented in dSPACE, consisted of the wind turbine control system. The control parameters and the hardware specications are included in Appendices B and D. The real-time experiment block diagram showing the hardware-in-the-loop implementation is shown in Fig. 10. The real-time experiment sequence is as follows: 1) The output of the three-mass model (i.e., the generator speed gen ) is an output from dSPACE as a reference to the motor drive (Unidrive). 2) The generator speed is measured through an incremental encoder and read back into dSPACE (introduction of measurement noise). 3) The measured speed (gen ) is used in the wind turbine control for the pitch controller, the optimal torque look-up table, and the torsional vibration damper.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
6 IEEE TRANSACTIONS ON ENERGY CONVERSION

Fig. 11.

LSS torque for Case 5: no uncertainty in mode frequencies. Fig. 13. LSS torque for Case 5: no uncertainty in mode frequencies.

Fig. 12.

LSS torque for Case 1: 10% uncertainty in F1 and +10% in F2 . Fig. 14. LSS torque for Case 1: 10% uncertainty in F1 and +10% in F2 .

4) The turbine controller in dSPACE outputs the resultant torque from the optimal and damping torques (gen , dam p ), as a torque reference (aero ) to the DSP. 5) The generator currents are measured through dSPACE to calculate the generator reaction torque. 6) The calculated torque is an input to the three-mass model to balance the aerodynamic torque aero and damp any torsional vibrations present. The grid-side converter control in the DSP is responsible for controlling the dc-link voltage and the reactive power ow to the grid. The experiment was set to output only active power by controlling the dc-link voltage. The reference for the dc-link voltage VDC was set from dSPACE. VI. RESULTS A. SimulationWind Step In order to compare the performance of the two torsional vibration dampers, the system shown in Fig. 1 was implemented in Simulink. A wind step from 14 to 24 m/s at 5 s was applied to the system to excite the system mode frequencies. A zero mean white Gaussian noise with a standard deviation of 0.5 rad/s was added to the generator speed to mimic measurement noise. The LSS torque was recorded in all simulations and used as a measure for comparing the performance of the two dampers. The simulation was performed for two cases outlined in Table I: Case 5 (no uncertainties in the mode frequencies) and Case 1 (worst case, with 10% uncertainty in the drive-train mode frequency F1 and +10% uncertainty in the blade in-plane symmetrical mode F2 ). Results are shown in Figs. 11 and 12. The simulation results for Case 5 are illustrated in Fig. 11. It can be observed that both dampers exhibited a good performance; however, the model-based damper response shows

Fig. 15.

Turbulent wind with a mean of 18 m/s.

fewer oscillations. The simulation results for Case 1 are shown in Fig. 12. It can be noted that the BPF-based torsional damper performance has been compromised since it failed to damp torsional vibrations effectively. In order to improve its performance, the BPF-based torsional damper had to be retuned. Nevertheless, the model-based torsional damper showed a good performance in both cases without any retuning needed. B. ExperimentWind Step A real-time hardware-in-the-loop experiment was carried out as described in Section V. The LSS torque data acquired from dSPACE were recorded and then plotted in MATLAB. The same simulation tests performed in the previous section were conducted. The LSS torque for Cases 5 and 1 is shown in Figs. 13 and 14, respectively. These results validate the simulation results shown in Figs. 11 and 12. C. ExperimentEffects of Turbulent Wind In reality, a wind turbine is exposed to turbulent winds and wind gusts rather than wind steps. Therefore, in order to test the performance of the two torsional vibration dampers in a realistic environment, the experiments were carried out again for a turbulent wind with a mean of 18 m/s as shown in Fig. 15.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
LICARI et al.: DAMPING OF TORSIONAL VIBRATIONS IN A VARIABLE-SPEED WIND TURBINE 7

designs. The downside of the proposed approach is that the design process is more challenging than for the conventional BPF-based damper. The model-based torsional damper can be implemented on different types of wind turbines provided that the converter topology used allows for control of the generator torque. APPENDIX A WIND TURBINE PARAMETERS
Fig. 16. LSS torque for Case 5: no uncertainty in mode frequencies.

Wind turbine: Power rating = 2 MW, rotor diameter = 38.75 m, rated speed = 18 r/min, blades = 3. Generator: Poles = 4, frequency = 50 Hz, stator resistance RS = 4.523 m, dq inductances Ld = Lq = 322 H, M = 1.75 Vs, time constant T = (Lq /RS ). Three-mass model: J1 = 2.227 106 kgm2 , J2 = 3.801 6 10 kgm2 , J3 = 416 633 kgm2 , K1 = 573.18 106 Nm/rad, K2 = 1.6 108 Nm/rad, D1 = D2 = 0 Nms/rad, N = 83:1. State-space representation of three-mass model 1 0 A= 0 0 0 K1 J1 0 K1 J2 0 0 1 J1 0 1 0 1 0 0 0 K2 J2 0 N K2 J3 0 0 0 1 N ;

Fig. 17.

LSS torque for Case 1: 10% uncertainty in F1 and +10% in F2 .

The LSS torque experimental results for Cases 5 and 1 are shown in Figs. 16 and 17, respectively. Experimental results for Case 5 show that the LSS torque amplitudes are slightly smaller for the case of the model-based vibration damper. The results for Case 1 show that the BPFbased vibration damper failed to damp torsional vibrations. This can be noted from the increase in the LSS torque amplitudes. On the other hand, the performance of the model-based vibration damper has hardly changed in the two cases. Moreover, its performance was superior to the BPF-based damper in both occasions. VII. CONCLUSION A performance comparison of two torsional vibration damper design approaches has been conducted through simulations and a real-time experiment. The results obtained are in good agreement. The stability analysis of the system with the two different dampers was assessed in the frequency domain. In the case of the BPF-based torsional damper, the analysis showed signicant deterioration in both phase and gain margins when model uncertainty was included. The stability margins of the system with the model-based damper were marginally affected. Simulation and experimental results showed that the performance of the model-based torsional damper was not affected in the presence of uncertainty whereas the BPF-based damper performance was compromised. Correspondingly, the BPF-based torsional damper had to be retuned to recover the intended performance. Due to its unaffected performance in the presence of model uncertainties, the model-based torsional damper eliminates retuning procedures associated with the conventional BPF-based

0 B= 0 0 0

0 0 ;C = [0 0 N2 J3

1];D = 0

x = [rot , (1 2 ), 2 , (2 3 ), gen ]T ; u = [ aero APPENDIX B CONTROLLERS PARAMETERS

m ]T .

Pitch controller: KP = 9.86 103 , KI = 3.4 103 , generator controller: KP = 0.136, KI = 25, grid controller: KP = 0.178, KI = 85. BPF-based damper: BPF1 : KBPF = 400, = 0.15, N = 15.07 rad/s, BPF2 : KBPF = 400, = 0.15, N = 24.5 rad/s, notch lter: 1 = 0.0015, 2 = 0.14, N = 11.31 rad/s. Model-based damper: R = 0.25, State feedback gain K = [6.4 104 ; 4.3 105 ; 1.4 103 ; 1.8 105 ; 6 102 ], Q = diag [2.56 1011 , 3 1015 , 6.22 1012 , 2.42 1011 , 1.649 103 ], notch: 1 = 0.0015, 2 = 0.14, N = 11.31 rad/s.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
8 IEEE TRANSACTIONS ON ENERGY CONVERSION

The model-based damper including the KF structure is shown in Fig. 18. Although the state-feedback controller and the KF have strong robustness properties when used individually, robustness is drastically reduced when they are used together. To mitigate this problem, the loop transfer recovery method was used to recover some of the robustness properties [34]. This was achieved by modifying the design of the KF. In this case, the plant uncertainties were represented by large amount of process noise as given by Qf = Q + q 2 Q (15)

where Q is the original process noise covariance matrix, q 2 Q is the additional ctitious noise matrix, and Qf is the resultant process noise covariance matrix. By adding ctitious process noise, the KF would interpret that there is less condence in the system model and hence would put more credibility on the measurements. Variable q was increased gradually to obtain the desired pole placement and an optimal KF. APPENDIX D HARDWARE SPECIFICATIONS
Fig. 18. loop. Detailed block diagram of the model-based torsional damper control

APPENDIX C MODEL-BASED DAMPER DESIGN PROCEDURE The design procedure of the model-based damper was divided in two parts. First, it was assumed that the states were available to design the state feedback controller and determine the state gain K. Then, a KF was designed to estimate the state variables. The measurement noise covariance R was obtained by taking ofine sample measurements. The standard deviation and the variance were then computed. The process noise covariance matrix Q was obtained by a systematic trial and error approach. The discrete-time KF equations assuming the measurement and process noise are stochastic, zero mean with known covariance matrices are given by x = F xk 1 + GUk k
Pk = F Pk 1 F T + Q Kk = Pk H T HPk H T + R 1

VSC: power rating = 10 kVA, C = 1020 F, switching frequency = 20 kHz, pulse-width modulation scheme = space vector pulse-width modulation. dSPACE model: DS1005, DSP model: TMS320F2808. Motor and generator: RPM = 3000 r/min, poles = 6, rated torque = 3.5 Nm, voltage = 400 V, power = 1.1 kW, encoder resolution = 4096 ppr. Variable-speed drive: power = 2.2 kW, voltage = 400 V. REFERENCES
[1] K. Z. Ostergaard, P. Brath, and J. Stoustrup, Gain-scheduled linear quadratic control of wind turbines operating at high wind speed, in Proc. IEEE Int. Conf. Control Appl., Oct. 2007, pp. 276281. [2] M. Molinas, J. A. Suul, and T. Undeland, Extending the life of gear box in wind generators by smoothing transient torque with statcom, IEEE Trans. Ind. Electron., vol. 57, no. 2, pp. 476484, Feb. 2010. [3] P. Cazelitz, W. Kleinkauf, T. Kruger, J. Petschenka, M. Reichard, and K. Storzel, Reduction of fatigue loads on wind energy converters by advanced control methods, in Proc. Eur. Wind Energy Conf., Dublin, Ireland, 1997. [4] V. Akhmatov, Variable-speed wind turbines with doubly-fed induction generators, part I: Modelling in dynamic simulation tools, Wind Eng., vol. 26, pp. 171188, 2002. [5] G. Ramtharan, Control of variable speed wind generators, PhD dissertation, The Univ. of Manchester, Manchester, U.K., 2008. [6] T. Burton, D. Sharpe, N. Jenkins, and E. A. Bossanyi, Wind Energy Handbook, 2nd ed. New York: Wiley, 2001, p. 498. [7] H. O. Rostoen, T. M. Undeland, and T. Gjengedal, Doubly-fed induction generator in a wind turbine, presented at the CIGRE Workshop Wind Power Impacts Power Syst., Oslo, Norway, 2002. [8] V. Akhmatov, Analysis of dynamic behaviour of electric power systems with large amount of wind power, PhD dissertation, Tech. Univ. of Denmark, Lyngby, Denmark, 2003. [9] V. Akhmatov, Modelling of variable-speed wind turbines with doubly-fed induction generators in short-term stability investigations, presented at the 3rd Int. Workshop Transmiss. Netw. Offshore Wind Farms, Stockholm, Sweden, 2002. [10] A. D. Wright and L. J. Fingersh, Part I: Control design, implementation, and initial tests, Nat. Renewable Energy Lab., Golden, CO, Tech. Rep. NREL/TP-500-42437 2008.

(10) (11) (12) (13) (14)

k k xk = x + Kk yk H x Pk = (I
Kk H) Pk

where F and G are discretized plant state and input matrices, xk 1 and Pk 1 are the initial conditions for the states and the estimation error covariance, Kk is the KF gain, Pk is the estimation error covariance matrix, H is the output matrix, Uk is the generator demanded torque, and yk is the measured speed [23]. The system was discretized using a sampling time of 20 ms.

This article has been accepted for inclusion in a future issue of this journal. Content is final as presented, with the exception of pagination.
LICARI et al.: DAMPING OF TORSIONAL VIBRATIONS IN A VARIABLE-SPEED WIND TURBINE 9

[11] M. Laubrock, T. P. Woldmann, and S. Bilges, Method for the active damping of the drive train in a wind energy plant, U.S. Patent 7501798B2, Mar. 10, 2009. [12] E. A. Bossanyi, The design of closed loop controllers for wind turbines, Wind Energy, vol. 3, pp. 149163, 2000. [13] E. A. Bossanyi, Wind turbine control for load reduction, Wind Energy, vol. 6, pp. 229244, 2003. [14] E. A. Bossanyi. (2009, Jun. 26). Controller for 5MW reference turbine [Online]. Available: www.upwind.eu/ [15] S. Suryanarayanan, A. Avangliano, and C. Barbu, Vibration damping method for variable speed wind turbines, U.S. Patent 7423352B2, Dec. 18, 2007. [16] J. Licari, C. E. Ugalde-Loo, J. Ekanayake, and N. Jenkins, Comparison of the performance of two torsional vibration dampers considering model uncertainties and parameter variation, presented at the Eur. Wind Energy Assoc. Annu. Meet., Copenhagen, Denmark, 2012. [17] J. Jun-Keun and S. Seung-Ki, Kalman lter and LQ based speed controller for torsional vibration suppression in a 2-mass motor drive system, IEEE Trans. Ind. Electron., vol. 42, no. 6, pp. 564571, Dec. 1995. [18] J. Licari, C. E. Ugalde-Loo, J. Liang, J. Ekanayake, and N. Jenkins, Torsional damping considering both shaft and blade exibilities, Wind Eng., vol. 36, pp. 181196, 2012. [19] H. Li and Z. Chen, Transient stability analysis of wind turbines with induction generators considering blades and shaft exibility, in Proc. 33rd Annu. Conf. IEEE Ind. Electron. Soc., Taiwan, Nov. 2007, pp. 1604 1609. [20] G. Ramtharan, N. Jenkins, O. Anaya-Lara, and E. Bossanyi, Inuence of rotor structural dynamics representations on the electrical transient performance of FSIG and DFIG wind turbines, Wind Energy, vol. 10, pp. 293301, 2007. [21] R. Krishnan, Permanent Magnet Synchronous and Brushless DC Motor Drives, 1st ed. New York: Taylor & Francis, 2010. [22] K. Ogata, Modern Control Engineering, 3rd ed. Englewood Cliffs, NJ: Prentice-Hall, 1997. [23] G. Welch and G. Bishop. (2001, Oct. 27). An Introduction to the Kalman Filter [Online]. Available: www.cs.unc.edu/welch/media/pdf/ kalman_intro.pdf [24] D. Simon, Optimal State Estimation, 1st ed. Mississauga, ON, Canada: Wiley, 2006. [25] B. Boukhezzar and H. Siguerdidjane, Nonlinear control of a variablespeed wind turbine using a two-mass model, IEEE Trans. Energy Convers., vol. 26, no. 1, pp. 149162, Mar. 2011. [26] D. Bourlis and J. A. M. Bleijs, A wind speed estimation method using adaptive Kalman ltering for a variable speed stall regulated wind turbine, in Proc. IEEE 11th Int. Conf. Probab. Methods Appl. Power Syst., Jun. 2010, pp. 8994. [27] K. Z. Ostergaard, P. Brath, and J. Stoustrup, Estimation of effective wind speed, J. Phys: Conf. Series, vol. 75, pp. 19, 2007. [28] N. Kodama, T. Matsuzaka, and N. Inomata, Power variation control of a wind turbine generator using probabilistic optimal control, including feed-forward control from wind speed, Wind Eng., vol. 24, pp. 1323, 2000. [29] J. R. Leigh, Control Theory, 1st ed. London, U.K.: Inst. Eng. Technol., 2004, vol. 64. [30] D. J. Leith and W. E. Leithead, Implementation of wind turbine controllers, Int. J. Control, vol. 66, pp. 349380, 1997. [31] J. F. Manwell, J. G. McGowan, and A. L. Rogers, Wind Energy Explained, 1st ed. New York: Wiley, 2002. [32] G. Hua, X. Dewei, W. Bin, and Y. Geng, Active damping for PMSGbased WECS with dc-link current estimation, IEEE Trans. Ind. Electron., vol. 58, no. 4, pp. 11101119, Apr. 2011. [33] H. Geng, D. Xu, B. Wu, and G. Yang, Comparison of oscillation damping capability in three power control strategies for PMSG-based WECS, Wind Energy, vol. 14, pp. 389406, 2011. [34] J. Doyle and G. Stein, Robustness with observers, IEEE Trans. Autom. Control, vol. AC-24, no. 4, pp. 607611, Aug. 1979.

John Licari (S09) was born in Malta, in 1980. He received the B.Eng. (Hons.) degree in electrical and electronics engineering from the University of Malta, Msida, Malta, in 2008. He is currently working towards his Ph.D. degree at the Institute of Energy, Cardiff University, Cardiff, U.K. His research interests include control of wind turbines, power electronic converters applied to renewable energy sources, power systems and energy efciency.

Carlos E. Ugalde-Loo (M02) was born in Mexico City, Mexico, in 1980. He received the B.Sc. degree in electronics and communications engineering from Tecnol gico de Monterrey, Monterrey, Mexico, the o M.Sc. degree in electrical engineering from Instituto Polit cnico Nacional, Mexico City, and the Ph.D. dee gree in electronics and electrical engineering from the University of Glasgow, Scotland, U.K., in 2002, 2005, and 2009, respectively. He is currently a Research Assistant at the Institute of Energy, Cardiff University, Cardiff, U.K. His research interests include power systems, FACTS devices, renewable energy generation and multivariable control.

Janaka B. Ekanayake (M95SM02) was born in Sri Lanka in 1964. He received the B.Sc. degree in electrical and electronic engineering from the University of Peradeniya, Peradeniya, Sri Lanka, and the Ph.D. degree in electrical engineering from the University of Manchester Institute of Science and Technology, Manchester, U.K., in 1990 and 1995, respectively. In 1992, he joined the University of Peradeniya, where he became a Professor of electrical and electronic engineering in 2003. He was also a Research Fellow at the University of Manchester. He joined the Cardiff University, Cardiff, U.K., as a Senior Lecturer in 2008. His main research interests include power electronic applications for power systems and renewable energy generation and its integration. Dr. Ekanayake is a Fellow of the Institute of Engineering and Technology.

Nicholas Jenkins (SM97F05) received the B.Sc. degree from Southampton University, Southampton, U.K., the M.Sc. degree from Reading University, Reading, U.K., and the Ph.D. degree from Imperial College London, London, U.K., in 1974, 1975, and 1986, respectively. He has 14 years of industrial experience, of which ve years were spent in developing countries. He was at the University of Manchester from 1992 to 1998. He is currently a Professor of renewable energy at Cardiff University, Cardiff, U.K. From 2009 to 2011, he was the Shimizu Visiting Professor at the Stanford University. Dr. Jenkins is the Director of the Institute of Energy, School of Engineering, Cardiff University. He is a Fellow of the Institute of Engineering and Technology and the Royal Academy of Engineering.

You might also like