Abstract
Wind turbine generators can be operated in various types of system configurations. Once the configurations change, the system oscillation mode shapes change accordingly. The modelling of the full system is necessary for studying this issue, yet it is quite hard. In this paper, a simple approach is developed to study the mode shapes of wind power systems without the necessity of adopting the complex full-system models. The key is that the q-d axis model of electric power system is transformed into the single-axis model, so that it could integrate with the equivalent circuit model of drive train mechanism. After analyzing some of the system configurations organized by the well-known MOD-2 wind turbine generator unit, the proposed approach is found to be effective for analyzing various oscillation modes such as the local torsional oscillations, as well as the inter-unit and inter-area oscillations.
THERE have been many investigations on the dynamics and transient stability of wind power systems. One of the important issues is about the oscillations of the drive train. The modelling of wind turbine generator (WTG) units plays an important role in these studies. In [
However, microgrids have recently become more and more popular, especially for the wind power systems and hybrid systems. The attentions are no longer paid on the oscillations of mechanical parts alone. It becomes significant to study the electromechanical oscillations, the inter-unit oscillations and the inter-area oscillations [
In order to study various oscillation modes of WTG systems in various system configurations, an approach opposed to the above-mentioned one is developed, where the mechanical systems are transformed and integrated into the electrical side. The key of the integration is that the generators (synchronous, induction or permanent magnet synchronous generators (PMSG)), converters (controlled rectifier and inverter) and circuits (transmission line, transformer, load and infinite bus) need to be transformed into the single-axis model, so that they can be combined with the wind turbine (WT) mechanical model. In this way, a WTG system, no matter how complex its configuration is, would turn to be an electrical circuit network. It will be possible to analyze such a network without difficulties, which is the merit of the proposed approach.
The WTG unit of MOD-2 system is adopted for investigation because complete studies of the unit are available [
For the WTG unit studied, the blade torque is transmitted to the generator via the drive train as shown in

Fig. 1 Drive train. (a) Scheme. (b) Mass-damping-spring model.

Fig. 2 Circuit model. (a) WTG unit. (b) Infinite/load bus.
For modelling a synchronous generator, the q-d axis model referred to the rotor reference frame is adopted, of which the electrical data are listed in
By combining the circuit model of the drive train and the simplified circuit model of the synchronous generator, the whole circuit model of WTG unit is obtained in
A grid can be viewed as a group of turbine generators operating in parallel. Thus, many circuit models of the turbine generator unit are combined in parallel to represent the circuit model of a grid. The result is shown in
A load bus can be represented the same as a grid, because it is also a power unit which does not generate but dissipate power. Therefore, the power flows are in the opposite direction. In addition, the loads usually have very small or zero inertia, and the shunt capacitance of the model will be very small. In case of an ideal constant-torque load, the model becomes a constant current source.
The MATLAB/Simulink/SimPowerSystems is used to analyze the wind power systems. Since the whole circuit model is adopted for the WTG unit, the wind power system including both the mechanical and the electrical parts can be built easily by using the electrical components. By simply using the built-in instructions, the state space equations can be obtained, and the eigenvalues and eigenvectors of the system can be further calculated.
Assume that the WTG unit operates at 1.0 p.u. of output power with power factor of 0.8. The turn ratio of the electromechanical interface transformer is thus equivalent to 1.8 by adopting 2.5 MW power base and 4.16 kV voltage base. The detailed calculation can be accessed in [
In order to verify the proposed approach, the oscillation modes for the WTG unit connecting to infinite bus (i.e., the WTG-I system) are studied. Firstly, the undamped and damped eigenvalues are calculated by ignoring and considering both the mechanical damping and the electrical resistance, respectively. The results are shown in
Compared with [

Fig. 3 Mode shapes of drive train of WTG.

Fig. 4 Mode shapes of WTG-I system.

Fig. 5 Mode shapes of WTG-L system.
In order to make a comparison between isolated and grid-connected systems, the case of the WTG-L system is further studied. It can be found that the generator mode and the EM mode present significant difference in shapes.
In
By comparing the WTG-L system with the WTG-I system, it can be found that the EM mode shapes of the two system configurations are totally different. For the WTG-I system configuration, the infinite bus ties together with the generator rotor, and the blades tie together with the hub. The combination of infinite bus and generator rotor forms a node, and the combination of blades and hub (i.e., WT) swings against it. Thus, the EM mode can also be named as the WT mode in the WTG-I system configuration. For the WTG-L system configuration, the combination of generator rotor, hub and blades forms a node, and the load bus swings against it. The EM mode can also be named as the load mode in the WTG-L system configuration. Therefore, the EM mode shape changes from the WT mode to the load mode when the configuration changes from the WTG-I system to the WTG-L system.
The mode shapes provide the basic information corresponding to mode behaviors. For the EM mode in the WTG-L system configuration, the combination of generator rotor, hub and blades forms a node, whereas a large magnitude of eigenvector appears at the load bus. This indicates that the EM mode is sensitive to electrical load disturbance, yet is totally insensitive to wind turbulence. From the viewpoint of control, it means that the dynamics of this mode can be improved by appropriate control of the system load, but cannot be improved by the pitch control or the generator automatic voltage regulator (AVR) control. However, for the EM mode in the WTG-I system configuration, a large magnitude of eigenvector appears at the combination of blades and hub. This indicates that the EM mode is sensitive to the turbulence of wind field, while the pitch control would be the effective way to provide damping for this mode. Therefore, there might be totally different characteristics under isolated and grid-connected wind power configurations.
To verify the different sensitivities subjected to wind turbulence for the WTG-L and WTG-I systems, the impulse responses excited by the blade torque disturbance for both systems are simulated. The hub-generator shaft torque and generator current are shown in

Fig. 6 Impulse responses excited by blade torque disturbance. (a) WTG-I system. (b) WTG-L system.
The same inference may be made for the generator mode. It can be easily judged from the shapes that both the generator AVR control and the electrical load have significant influences on the generator mode behaviors in the WTG-L system configuration. However, only the generator AVR control is important in the WTG-I system configuration.
Moreover, the changes in mode frequency can be a serious problem in some cases. For some WTG units with torsional vibration problems in main shaft, the frequency-based countermeasures may be adopted to alleviate the vibrations. The vibration suppression effect may be worse or even totally futile if the mode frequency is changed.
Typically, a WTG unit is connected to network via a step-up transformer and a transmission line. The combined impedance of transformer and transmission line is found to significantly affect the mode shapes of the wind power system. This effect is studied for the WTG-I system. It is assumed that the combined impedances of short and long transmission lines are the same as and 3000 times that of the generator, respectively.
Based on the simulation results, it is found that there are obvious changes in the EM mode shapes. This is shown in

Fig. 7 EM mode shapes for different lengths of transmission lines. (a) Long transmission line. (b) Short transmission line.
The shape variation process for the EM mode is analyzed and shown in

Fig. 8 Variation process of EM mode shapes. (a) Shape variation process. (b) Frequency variation with K factor.
In addition to the isolated and grid-connected system configurations, the WTG units may be clustered to form a microgrid. The proposed approach can be applied to such a multi-machine system without any difficulty.
For the system of two paralleled WTG units supplying power to loads (i.e., the two-WTG-to-load system), the vibration modes are shown in

Fig. 9 Mode shapes of two-WTG-to-load system.
Since the mechanical part shapes do not change with the electrical power system configurations, the blades and hub sections are combined and simplified as a WT section in order to simplify the mode shape diagrams of a multi-machine system. The mode shapes of a two-WTG-interconnected system for different lengths of transmission lines are shown in

Fig. 10 Mode shapes of two-WTG-interconnected system.
In the short transmission line situation, it can be found that the mode shapes are the same as those of the two-WTG-to-load system, except that a new load mode is produced, in which two load buses swing to each other. When the transmission line becomes long, two changes happen. One is that for the WT mode and one of the generator modes, the originally united load buses become swing to each other. The other is that for the WT mode, the generator rotor separates from the node formed by the generator rotors and loads, and turns to tie together with the WT. For a three-WTG-interconnected system, the mode shapes are shown in Fig. A2 of Appendix A.
According to the equivalent circuit of WTG unit, the turn ratio of interface transformer depends on the generator operation conditions. Therefore, it is evident that some of the mode frequencies will vary following the variation in generator operation conditions. In case of the power output keeping at 1.0 p.u., the frequencies of generator mode and EM mode are calculated and listed in
One of the merits of the proposed WTG unit model is that the poles of a complex multi-machine system can be easily derived. Thus, the dynamics of each mode can be assessed. The WTG-I system is studied as an example, and three cases are analyzed.
Referring to Figs.

Fig. 11 Effects on pole locus. (a) Damping of hub-generator shaft. (b) Generator resistance. (c) Transmission line impedance.
For the generator mode ( rad/s), the locus is away from the vertical axis, which means that the mode gets more stable. For the hub mode ( rad/s) and the EM mode ( rad/s), their tendency is similar. Initially, the locus is toward the horizon axis and away from the vertical axis. Hence, the effect of DHG is also positive. The dynamics of the agonizing EM mode, of which the decrement factor is only -0.0234, can also be improved by a damper installed at the hub-generator shaft section. When the damping keeps increasing, the complex number pole pair turns to become the real number poles. If the damping increases even larger, new complex number pole pair will be generated. However, the locus will move away from the horizon axis and toward the vertical axis. That means the increase in damping will have a negative effect in this situation.
The generator resistance will affect the modes in relevance to electrical parts. This is shown in
The effect of transmission line impedance is shown in
Another merit of the proposed WTG unit model is that the Bode plots and impulse responses can be easily obtained for a complex wind power system. The responses due to the interactions between the electrical and the mechanical parts can thus be assessed. As an example, the WTG-I system is studied. The results are shown in

Fig. 12 Frequency responses. (a) Excited by generator voltage disturbance. (b) Excited by blade torque disturbance.
The responses shown in
The responses shown in
By the cross comparison between

Fig. 13 Impulse responses. (a) Excited by generator voltage disturbance. (b) Excited by blade torque disturbance.
The proposed WTG unit model can also be used to assess the interactions between two WTG units in a multi-machine system. The two-WTG-to-load system is studied as an example. The resulting Bode plots are shown in

Fig. 14 Frequency responses of two-WTG-to-load system.
Although the MOD-2 adopts a synchronous generator, most of commercial WTGs are designed using squirrel cage induction generators. It is interesting to know what will be different if the synchronous generator is replaced by an induction generator.
In [
A direct drive WTG unit includes converters as the grid interfaces. The typical scheme is shown in

Fig. 15 Diagram of direct drive WTG system.

Fig. 16 Complete model of direct drive WTG unit.
However, the modelling of converters is based on the presumption of constant control instruction. When it comes to the emerging sub-synchronous oscillation phenomenon [
Compared with the full machine model, the simplified machine model has neglected the transients of stator and rotor windings, and the effects of damping windings and salient pole. These simplifications result in significant deviations in transient responses. The transient responses of the machine rotor angle are simulated for the WTG-I system configuration with the simplified model, full model, and full model plus power system stabilizer (PSS), respectively. A three-phase-to-ground fault occurs at 1.0 s and disappears at 1.05 s at the far end of transmission line. Different mechanical damping as well as different transmission line lengths have been considered in the simulations. The overall results are shown in

Fig. 17 Rotor angle responses. (a) Decreased mechanical damping/short transmission line. (b) Decreased mechanical damping/long transmission line. (c) Increased mechanical damping/short transmission line. (d) Increased mechanical damping/long transmission line.
Furthermore, by comparing the effects of mechanical damping and transmission line length, it can be seen that the increased mechanical damping and the long transmission line will be helpful for decreasing the error. Therefore, the accuracy is dependent on the application situations.
The main contribution of the proposed approach is that not only the local torsional oscillation modes, but also the electromechanical, the inter-unit and even the inter-area modes can be studied without the necessity to adopt the complex full model.
It is easily applied to solve the commonly encountered sub-synchronous resonance and oscillation problems to find the torsional interaction or transient amplification effects. However, the capability is limited because only a simplified machine model is adopted. The lack of control systems will make it hard to study the emerging sub-synchronous oscillation phenomenon. Nevertheless, this can be set as a research goal for future modification and improvement.
After using the proposed approach to study some of the different configurations of WTG systems, the following noteworthy results are summarized.
1) The electrical systems would affect the shapes of some of the oscillation modes, even the mechanical modes.
2) The EM mode shape is significantly dependent on system configurations.
3) The transmission line length is an important factor to influence the mode shapes of WTG systems.
Appendix
The mode shapes of three-WTG-to-load and three-WTG-interconnected systems are shown in Figs. A1 and A2, respectively.

Fig. A1 Mode shapes of three-WTG-to-load system.

Fig. A2 Mode shapes of three-WTG-interconnected system.
For a PMSG, the typical dynamic equations in q-d reference frame are as the following formulas.
(B1) |
where P is the pole number; p is the differentiation operator; is the d-axis voltage; is the d-axis current; is the q-axis voltage; is the q-axis current; is the stator resistance; is the d-axis stator inductance; is the q-axis stator inductance; is the q-d axis flux linkage; is the rotor angular velocity (electrical); and is the EM torque.
Assume that the initial angle of rotor reference frame is zero degree, thus , and the relationship between the d-axis and q-axis currents becomes:
(B2) |
Rewriting the q-axis equation by using the above formula, we have:
(B3) |
Usually, , and the term can be neglected. Moreover, by changing the electrical angular velocity to mechanical angular velocity and the q-d axis flux linkage to the abc axis flux linkage , the equations can be obtained as the followings. Based on those equations, the single-axis circuit model can be obtained as shown in Fig. B1.
(B4) |

Fig. B1 Single-axis model of PMSG.
The rectifier scheme is shown in Fig. B2(a). Assume that the input voltages of the rectifier are ideal sinusoidal waves with peak value and angular velocity . The phase-a voltage can be expressed as:
(B5) |

Fig. B2 Rectifier. (a) Scheme. (b) Ideal transformer model.
By using the Park’s transformation to the three-phase voltages , and , we obtain:
(B6) |
Assume the delay angle of the rectifier to be , and neglect the commutation loss. Then the output voltage of the rectifier will be:
(B7) |
The power invariant is applied to the rectifier assuming that no losses are considered. So, the rectifier can be modeled as an ideal transformer as shown in Fig. B2(b).
The pulse width modulation (PWM) inverter scheme studied is shown in Fig. B3(a). Its operation can be analyzed by expressing the output voltage of the inverter as a product of the six-step continuous phase voltage and a pulse train with pulses of unit magnitude. For example, , the voltage of phase a, may be expressed as the following formula.

Fig. B3 PWM inverter. (a) Scheme. (b) Ideal transformer model.
(B8) |
where is the input voltage of the inverter; and is the pulse train, which may be expressed in Fourier series as:
(B9) |
where ; and .
By using the Park’s transformation to the voltages , and , we obtain:
(B10) |
If the harmonics are neglected, the voltage formulas are further reduced to:
(B11) |
Again, the power invariant applies to the inverter if it is assumed that no losses are considered. So, the inverter can be modeled as an ideal transformer, as shown in Fig. B3(b).
REFERENCES
S. K. Kim and E. S. Kim, “PSCAD/EMTDC-based modeling and analysis of a gearless variable speed wind turbine,” IEEE Transactions on Energy Conversion, vol. 22, no. 2, pp. 421-430, Jun. 2007. [百度学术]
S. K. Salman and A. L. J. Teo, “Windmill modeling consideration and factors influencing the stability of a grid-connected wind power-based embedded generator,” IEEE Transactions Power Systems, vol. 18, no. 2, pp. 793-802, May 2003. [百度学术]
H. Li and Z. Chen, “Transient stability analysis of wind turbines with induction generators considering blades and shaft flexibility,” in Proceedings of the 33rd Annual Conference of the IEEE Industrial Electronics Society (IECON), Taipei, China, Nov. 2007, pp. 1604-1609. [百度学术]
S. M. Muyeen, M. H. Ali, R. Takahashi et al., “Comparative study on transient stability analysis of wind turbine generator system using different drive train models,” IET Renewable Power Generation, vol. 1, no. 2, pp. 131-141, Jun. 2007. [百度学术]
J. Peeters, D. Vandepitte, and P. Sas, “Analysis of internal drive train dynamics in a wind turbine,” Wind Energy, vol. 9, no. 1-2, pp. 141-161, Jun. 2006. [百度学术]
O. Lundvall, N. Stromberg, and A. Klarbring, “A flexible multi-body approach for frictional contact in spur gears,” Journal of Sound and Vibration, vol. 278, no. 3, pp. 479-499, Dec. 2004. [百度学术]
M. Jafarian and A. M. Ranjbar, “Interaction of the dynamics of doubly fed wind generators with power system electromechanical oscillations,” IET Renewable Power Generation, vol. 7, no. 2, pp. 89-97, Mar. 2013. [百度学术]
E. Rezaei, A. Tabesh, and M. Ebrahimi, “Dynamic model and control of DFIG wind energy systems based on power transfer matrix,” IEEE Transactions on Power Delivery, vol. 27, no. 3, pp. 1485-1493, Jul. 2012. [百度学术]
L. Fan, H. Yin, and Z. Miao, “On active/reactive power modulation of DFIG-based wind generation for interarea oscillation damping,” IEEE Transactions on Energy Conversion, vol. 26, no. 2, pp. 513-521, Jun. 2011. [百度学术]
L. Wang and K. H. Wang, “Dynamic stability analysis of a DFIG-based offshore wind farm connected to a power grid through an HVDC link,” IEEE Transactions on Power Systems, vol. 26, no. 3, pp. 1501-1510, Aug. 2011. [百度学术]
A. J. Laguna, N. F. B. Diepeveen, and J. W. V. Wingerden, “Analysis of dynamics of fluid power drive-trains for variable speed wind turbines: parameter study,” IET Renewable Power Generation, vol. 8, no. 4, pp. 398-410, May 2014. [百度学术]
J. A. M. Bleijs, “Wind turbine dynamic response-difference between connection to large utility network and isolated diesel micro-grid,” IET Renewable Power Generation, vol. 1, no. 2, pp. 95-106, Jun. 2007. [百度学术]
O. Wasynczuk, D. T. Man, and J. P. Sullivan, “Dynamic behavior of a class of wind turbine generators during random wind fluctuation,” IEEE Transactions on Power Apparatus and Systems, vol. 100, no. 6, pp. 2837-2845, Jun. 1981. [百度学术]
P. C. Krause and O. Wasynczuk, “Method of resynchronizing wind turbine generators,” IEEE Transactions on Power Apparatus and Systems, vol. 100, no. 10, pp. 4301-4308, Oct. 1981. [百度学术]
R. Wamkeue, Y. Chrourou, J. Song et al., “State modeling based prediction of torsional resonances for horizontal-axis drive train wind turbine,” in Proceedings of IET Conference on Renewable Power Generation (RPG 2011), Edinburgh, UK, Sept. 2011, pp. 1-5. [百度学术]
C. H. Lin, “Modeling and analysis of power trains for hybrid electric vehicles,” International Transactions on Electrical Energy Systems, vol. 28, no. 6, pp. 1-19, Jun. 2018. [百度学术]
W. M. Lin, C. C. Tsai, and C. H. Lin, “Analysing the linear equivalent circuits of electromechanical systems for steam turbine generator units,” IET Generation, Transmission & Distribution, vol. 5, no. 7, pp. 685-693, Jul. 2011. [百度学术]
D. Shu, X. Xie, H. Rao et al., “Sub- and super-synchronous interactions between STATCOMs and weak AC/DC transmissions with series compensations,” IEEE Transactions on Power Electronics, vol. 33, no. 9, pp. 7424-7437, Sept. 2018. [百度学术]
Y. Xu, H. Nian, T. Wang et al., “Frequency coupling characteristic modeling and stability analysis of doubly fed induction generator,” IEEE Transactions on Energy Conversion, vol. 33, no. 3, pp. 1475-1486, Jan. 2018. [百度学术]