Vous êtes sur la page 1sur 6

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 15, NO.

4, NOVEMBER 2000

1287

A Comparison of Classical, Robust, and Decentralized Control Designs for Multiple Power System Stabilizers
George E. Boukarim, Shaopeng Wang, Joe H. Chow, Fellow, IEEE, Glauco N. Taranto, Member, IEEE, and Nelson Martins, Fellow, IEEE

AbstractSeveral control design techniques, namely, the classical phase compensation approach, the -synthesis, and a linear matrix inequality technique, are used to coordinate two power system stabilizers to stabilize a 5-machine equivalent of the South/Southeast Brazilian system. The open-loop system has an unstable intearea mode and cannot be stabilized using only one conventional power system stabilizer. Both centralized and decentralized controllers are considered. The different designs are compared and several interesting observations are provided. Index TermsDecentralized control, linear matrix inequalities, power system stabilizers, robust control, small-signal stability.
Fig. 1. Study system configuration.

I. INTRODUCTION HE SECURE operation of power systems requires the application of robust controllers, such as Power System Stabilizers (PSS), to provide sufficient damping at all credible operating conditions [1]. Recently, many researchers have invesoptigated the use of robust control techniques including timization [2] and -synthesis [3] for developing advanced and automated procedures for power system damping controller design. A partial list of these results is contained in [4][8]. In this paper, we compare several control techniques for the design of two PSSs in a 5-machine equivalent of the South/Southeast Brazilian System. This system needs at least two conventional PSSs to operate in a stable manner, and thus it is an ideal example to illustrate advanced control design techniques that can coordinate more than one control loop. The control design techniques evaluated include a synthesis design utilizing a structured uncertainty model which captures the variation of the operating conditions, and a Linear Matrix Inequality (LMI) approach in which low-order PSSs are designed directly. The LMI approach also allows for a decentralized design. As a baseline comparison, we also include a design using the classical phase compensation approach. Thus the comparisons span the spectra of high-order versus low-order, centralized versus decentralized, and coordinated versus uncoordinated controllers. Due to the space limitation, we only provide an
Manuscript received April 27, 1999. The work in this paper was supported in part by an EPRI/DOD grant, administrated through Carnegie Mellon University. G. E. Boukarim, S. Wang, and J. H. Chow are with the ECSE & EPE Department, Rensselaer Polytechnic Institute, Troy, NY 12180-3590 USA. G. N. Taranto is with the Electrical Engineering Department, COPPE/UFRJ, Rio de Janeiro, RJ 21945-970, Brazil. N. Martins is with CEPEL, Rio de Janeiro, RJ 21944-970, Brazil. Publisher Item Identifier S 0885-8950(00)10361-X.

overview of the design methods. The formulas used for the design are not shown in the paper. However, references are provided for readers interested in further pursuing the methods. The paper is organized as follows: Section II provides a discussion of the 5-machine Brazilian system. Section III contains an overview of the design methods used in this paper and shows the designed controllers. The performance of the controllers for a system disturbance and the comparison of the controllers are contained in Section IV. II. TEST SYSTEM The test system is a modified 7-bus, 5-machine equivalent model of the South/Southeast Brazilian system first presented in [10] and depicted in Fig. 1. The complete system data can be obtained from [10]. The modal analysis of the system indicates that there are two interarea modes. Mode 1 is due to the Southeast (SE) equivalent system oscillating against the Itaipu generator, while Mode 2 is due to the South system (represented by Santiago, Segredo and Areia) oscillating against the Southeast system together with the Itaipu generator. The system also has two local modes of oscillations within the South system: Mode 3 consisting of Areia and Segredo oscillating against Santiago, and Mode 4 consisting of Areia oscillating against Segredo. Table I presents five operating scenarios, obtained from the same transmission network configuration, where the values and (connecting Buses 56 and of the reactances Buses 67, respectively) are varied. The generator and load levels are the same for the 5 scenarios. Table I also shows the frequencies and the damping ratios of Modes 1 and 2. In all operating scenarios, Mode 1 is open-loop unstable. The local modes (Modes 3 and 4) have damping ratios of about 20% and oscillation frequency of about 1.46 Hz.

08858950/00$10.00 2000 IEEE

1288

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 15, NO. 4, NOVEMBER 2000

TABLE I OPEN-LOOP OPERATING CONDITIONS

TABLE II CLOSED-LOOP SYSTEM DAMPING RATIO (%)

Fig. 3.

Uncertainty model formulation.

A. Classical Design
Fig. 2. Pole-zero locations.

This open-loop unstable system presents a control design challenge, as attempts to stabilize the system applying only one PSS in the form of a conventional two-section lead compensation fail. The reason for this limitation is due to the presence of unstable right-half-plane (RHP) zeros that occur in the single-input single-output (SISO) transfer function from the AVR reference voltage to the machine rotor speed regardless of the generator considered. For example, in Scenario 5, for the transfer function from the Itaipu generator AVR to the machine , and for the transfer rotor speed, the zeros occur at function from the Segredo generator AVR to the machine rotor , where both sets of zeros speed, the zeros occur at . The solution are close to the unstable poles at to this problem is to apply two PSSs, one of which must be located at the Itaipu generator. Fig. 2 shows the pole-zero locations (Case 5) for the transfer function from the Itaipu and Segredo AVRs to the machine rotor speeds. Note that the system no longer has RHP zeros, thus allowing the stabilization of the system using conventional PSSs. Also note that Mode 4 is very close to a zero, and thus it is not affected by the controller. The control design will influence Mode 3, which will be discussed later.

In the classical design, a double-lead compensator is applied individually to each of the Segredo and Itaipu generators excitation systems. Each of the compensators is designed individually with the second compensator loop remaining open, and with the objective of providing maximum damping benefit to modes 1 and 2 for all 5 operating conditions. Note that this is an uncoordinated design and that neither compensator can achieve stability by itself due to the presence of the RHP zeros in the open-loop transfer function of both compensators. Stability is achieved however, when both compensators are applied simultaneously as discussed in the previous section. The designed compensators, including a washout stage, have the following transfer functions: Segredo PSS PSS Itaipu PSS PSS (2) (1)

III. DESIGN APPROACHES AND RESULTS In this section we describe the procedures used for designing the Segredo and Itaipu PSSs. We overview the design methods and present the design results. Interested readers may refer to the cited literature for further details. All the control designs are performed on models linearized about the 5 operating scenarios. To summarize the design results, the two modes with the smallest damping ratios for each design will be listed in Table II.

The combination of these two PSSs provides satisfactorily damping to Mode 1 ranging from 7.2% to 10.6% (Table II). Modes 2 has at least 50% damping. However, as common in the classical design, the PSSs interact with the machine dynamics and excitation systems to produce a lightly damped control mode (mode x in Table II). This classical design will be referred to as Design C. The frequency responses of these PSSs without the washout stages are shown in Figs. 4 and 5. B. -Synthesis Design

The design model in the -synthesis framework is shown in is a nominal system model, and Fig. 3. In this form, are the inputs to the AVR summing junction of the Segredo and are the machine and Itaipu machines, respectively,

BOUKARIM et al.: A COMPARISON OF CLASSICAL, ROBUST, AND DECENTRALIZED CONTROL DESIGNS

1289

Fig. 4. Frequency response for PSS , designs C and LD.

Fig. 6.

Frequency response for designs M and LC.

spectrum, weighting functions are applied to the input and output vectors in the design model. The weighting function

Fig. 5. Frequency response for PSS , designs C and LD.

(3)

speeds, and is the two-input, two-output PSS to be deand take on signed. Given that the line impedances different values for the different operating scenarios, they are considered uncertain, and so is the actual system model which is a function of these impedances. The matrix, known as the uncertainty matrix, accounts for the difference between the and the actual system model. For nominal system model produces a difeach value of , the closed loop of and ferent system model, and as takes on its different values the closed-loop combination generates the actual system model for all the operating scenarios. The uncertainty matrix , the nom, and the manner in which they are coninal system model nected to generate the actual system model are referred to as the uncertainty model. The vectors and provide this connection and , and reflect how the uncertainty matrix between acts to modify the nominal system dynamics to match those of the actual system. Procedures to develop the uncertainty model can be found in [12] and [8]. In these procedures, the linearized system state matrices at the various operating conditions are used to obtain a representation of these matrices as matrix polynomials in the uncertain parameters. These matrix polynomials are then represented in the form of Linear Fractional Transformations (LFTs) and combined to yield the system model shown in Fig. 3. The matrix in this case is a real diagonal matrix with and , along its dithe uncertain parameters, namely . The nominal model agonal, and has the dimensions of is a product of the polynomial and LFT representations and is, in general, different from the system model at any of the operating scenarios used to generate the uncertainty model. To influence the -synthesis algorithm and produce controllers with desirable characteristics throughout the frequency

, and to the output is applied to the input vector, , of the nominal system. vector, The design model, including the weighting functions, has 47 states and is reduced to 21 states using the Hankel norm approximation [15] prior to the controller design. A 21-state controller is then obtained and reduced to the 8th-order controller given in the Appendix. The washout filters are included in the controllers afterwards. This design will be referred to as Design M, and the damping ratios for Modes 1 and 2 achieved with this controller are shown in Table II. The frequency response of this controller is shown in Fig. 6. This centralized control is a wide-area stability control [13] which requires the communication of measurements that are not locally available. For the test system, the two generators are about 200 miles apart. Thus the implementation of this control strategy would be more expensive than those utilizing only local measurements. In addition, communication delays reduce the performance of the controller. When the delay is less than 35 msec, the closed-loop systems for all 5 operating conditions are still stable but have less damping. It should be mentioned that this controller cannot stabilize the systems if the off-diagonal terms are lost. C. LMI Designs The Linear Matrix Inequality approach presented in [14], [16] is used to design a low-order centralized and a low-order decentralized controller. In an earlier paper [14], the technique was used to design a PSS to stabilize a single-machine infinite-bus system at 5 operating scenarios. Here we are extending the results to a two-PSS design.

1290

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 15, NO. 4, NOVEMBER 2000

It is well-known that the LMI approach can solve robust control design problems with constraints such as regional poleplacement [17] through a convex optimization algorithm [18]. However, the controller obtained is of full order, that is, the same size as the design model including weighting functions. When the design problem is formulated for a low-order controller, the resulting optimization is nonconvex [19]. To circumvent the nonconvexity, the low-order controller LMI approach in [14] uses coprime factorization. The coprime factors allow the control design to be formulated as an approximate pole-placement norm of a stable transfer funcproblem of minimizing the tion. The minimization can be extended to include more than one operating condition. The stability of the closed-loop system can be guaranteed by imposing positive realness conditions. This LMI approach can readily handle decentralized control by simply enforcing the gains in the off-diagonal entries of the controller to be zero in the convex optimization. Using this approach by adjusting the weights to achieve about the same Mode 1 damping as Design M, the centralized controller obtained is

an iterative LMI algorithm is used in [20] to design the decentralized PSSs. However, the noniterative method used here and in [14] is more efficient. Table II shows that Design LD achieves about the same damping improvement for Mode 1 as Design LC, but with less damping reduction of Mode 3. The frequency responses of these two controllers are shown in Figs. 4 and 5. Note that although the frequency responses of Design C and LD are quite similar, their effects on the closed-loop system poles are very different. D. Comparisons Based on Small-Signal Stability We provide two different comparisons of the four control designs. The first comparison, presented in this subsection, is based on the damping analysis of the major modes of oscillation as given by Table II, and the second one, presented in the next section, is based on time response simulations. Based on the damping ratio enhancements noted in Table II, as compared to Table I, we arrive at the following conclusions: 1) All four designs provide controllers that enhance the damping of Modes 1 and 2 in all 5 operating scenarios. 2) The 4th-order centralized Design LC matches the performance of the 8th-order centralized Design M. 3) The coordinated decentralized Design LD provides better damping enhancement of Mode 1 over all 5 operating scenarios than the classical decentralized Design C. 4) From the frequency response plots, the centralized Designs M and LC achieve the same damping enhancement with much smaller gains (an order of magnitude) than the decentralized design. This illustrates the potential benefit in control design when more information is available. 5) The decentralized controllers are phase-lead compensators, whereas the phase characteristics of the centralized controllers are more difficult to assess and to explain. IV. STEP RESPONSE SIMULATION COMPARISON
AND-CONTROLLER

(4) in each channel is Note that the washout filter not shown in the controller. The poles of the controller are and , and the zeros are at at , and 2.1. It should be pointed out that, even if the off-diagonal controllers are lost, the diagonal control terms can still stabilize the system for all 5 operating scenarios but with much less damping. When the communication delays are less than 18 msec, the closed-loop systems remain stable. The centralized LMI controller is denoted as Design LC. The frequency response of this controller is also shown in Fig. 6. As shown in Table II, this design results in substantially more damping for Mode 1 than Designs C and M. Mode 2 has at least 31.6% damping, but the damping on Mode 3 has been reduced. However, the Mode 3 damping reduction can be readily remedied by reducing the controller gain, because Mode 1 has more than sufficient damping. The decentralized controller obtained, again by adjusting the weights to achieve about the same Mode 1 damping as Design M, from the LMI design is Segredo PSS PSS Itaipu PSS PSS (6) (5)

Although not having only real poles and zeros as in conventional PSS, these PSSs are also phase-lead compensators. The decentralized LMI controller is denoted as Design LD. Note that

To further test the performance of the controllers, a disturbance is applied to the closed-loop systems using the PacDyn Program [10]. The disturbance is a simultaneous application of a positive perturbation in the mechanical power of the Itaipu generator and a negative perturbation in the mechanical power of the SE equivalent machine. The resulting Itaipu rotor speed transients are shown in Fig. 7. From the simulations, we arrive at some further conclusions: 1) The speed variations are higher for the centralized controllers than for the decentralized controllers. This is to be expected because the decentralized controllers have higher gains and hence better disturbance rejection capability for rotor speeds. 2) The centralized Design M results in more transients than the Design LC, because a higher-order controller has more internal transients that also need to decay. 3) The low-order centralized Design LC exhibits mostly a single modal response.

BOUKARIM et al.: A COMPARISON OF CLASSICAL, ROBUST, AND DECENTRALIZED CONTROL DESIGNS

1291

REFERENCES
[1] J. F. Hauer, Robust damping control for large power systems, IEEE Control Systems Magazine, vol. 9, no. 1, pp. 1218, 1989. [2] J. C. Doyle, K. Glover, P. Khargonekar, and B. A. Francis, State-space solutions to standard H and H control problems, IEEE Trans. Automatic Control, vol. 34, pp. 831847, 1989. [3] G. J. Balas, J. C. Doyle, K. Glover, A. Packard, and R. Smith, -Analysis and Synthesis Toolbox: The MathWorks, Inc., 1995. [4] M. Klein, L. X. Le, G. J. Rogers, S. Farrokpay, and N. J. Balu, H damping controller design in large power system, IEEE Trans. Power Systems, vol. 10, pp. 158166, 1995. [5] G. N. Taranto and J. H. Chow, A robust frequency domain optimization technique for mining series compensation damping controllers, IEEE Trans. Power Systems, vol. 10, no. 3, pp. 12191225, 1995. [6] Q. Zhao and J. Jiang, Robust SVC controller design for improving power system damping, IEEE Trans. Power Systems, vol. 10, pp. 19271932, 1995. [7] S. Chen and O. P. Malik, H optimization based power system stabilizer design, IEE Proc., Pt. C, pp. 179184, 1995. [8] M. Djukanovic, M. Khammash, and V. Vittal, Application of the structured singular value theory for robust stability and control analysis in multimachine power systemsPart I: Framework development, in 1998 IEEE Winter Power Meeting, Tempa, FL. [9] , Application of the structured singular value theory for robust stability and control analysis in multimachine power systemsPart II: Numerical simulations and results, in 1998 IEEE Winter Power Meeting, Tempa, FL. [10] N. Martins and L. T. G. Lima, Eigenvalue and frequency domain analysis of small-signal electromechanical stability problems, IEEE Special Publication on Eigenanalysis and Frequency Domain Methods for System Dynamic Performance, pp. 1733, 1989. [11] E. V. Larsen and D. A. Swann, Applying power system stabilizersParts IIII, IEEE Trans. Power Apparatus and Systems, vol. PAS-100, pp. 30173046, 1981. [12] G. E. Boukarim and J. H. Chow, Modeling of nonlinear system uncertainties using a linear fractional transformation approach, in Proc. 1998 American Control Conference, pp. 29732979. [13] I. Kamwa, L. Gerin-Lajoie, and G. Trudel, Multi-loop power system stabilizers using wide-area synchronous phasor measurement, in Proc. 1998 American Control Conference, pp. 29632967. [14] S. Wang and J. H. Chow, Coprime factors, strictly positive real functions, LMI and low-order controller design for SISO systems, in Proc. 1998 Conference on Decision and Control, pp. 27382744. [15] K. Zhou, Essentials of Robust Control: Prentice-Hall Inc., 1998. [16] S. Wang and J. H. Chow, Low-order controller design for model matching optimization using coprime factors and linear matrix inequalities, in Proc. 1999 American Control Conference. [17] M. Chilali and P. Gahinet, H design with pole placement constraints: An LMI approach, IEEE Trans. on Automatic Control, vol. 41, pp. 358367, 1996. [18] P. Gahinet, A. Nemirovskii, A. J. Laub, and M. Chilali, LMI Control Toolbox: For Use with MATLAB: The MathWorks, Inc., 1995.

Fig. 7.

Disturbance simulation of controllers.

4) The classical Design C shows additional initial transients due to the lightly damped control mode. The optimized decentralized Design LD shows smaller initial transients. V. CONCLUSION This paper examines several methods of designing multiple PSSs for damping power system swings. A model of the Brazilian system requiring more than two PSSs for stabilization is used as the test system. It is found that both decentralized and centralized controllers provide good damping enhancement to the interarea modes. The centralized controllers require much low gain to achieve the same amount of damping enhancement than the decentralized controllers. However, the centralized controllers have less disturbance rejection capabilities and require fast communication links to implement. APPENDIX The controller matrices of Design M are in the equation at the top of the page.

1292

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 15, NO. 4, NOVEMBER 2000

[19] C. Scherer, P. Gahinet, and M. Chilali, Multiobjective output-feedback control via LMI optimization, IEEE Transactions on Automatic Control, vol. 42, pp. 896911, 1997. [20] G. N. Taranto, S. Wang, J. H. Chow, and N. Martins, Decentralized design of power system damping controllers using a linear matrix inequality algorithm, in Proc. of VI SE-POPE, 1998.

Joe H. Chow received his B.S. degrees in electrical engineering and mathematics from the University of Minnesota, Minneapolis, and his M.S. and Ph.D. degrees in electrical engineering from the University of Illinois, Urbana. From 1978 to 1987, he was with GE Power Systems working on power system dynamics and control design. He is currently a Professor of Electrical, Computer, and Systems Engineering, and Electric Power Engineering at Rensselaer Polytechnic Institute. His interests include multivariable control systems, large-scale systems, and power system dynamics and control.

George E. Boukarim received his B.S., M.E., and Ph.D. degrees in 1987, 1988, and 1998 respectively, all in electric power engineering, from Rensselaer Polytechnic Institute in Troy, NY. From 1988 to 1994 and from 1998 to present he was with the Power Systems Energy Consulting (PSEC) Department of the General Electric Company working in the area of power system dynamics and control. He also worked for the Transmission Technology Institute, ABB Power T&D Company, for a short period in 19971998. His interest include robust multivariable control, and power system dynamics and control.

Glauco N. Taranto received his B.Sc. (1988) from the State University of Rio de Janeiro, M.Sc. (1991) from the Catholic University of Rio de Janeiro, and Ph.D. (1994) from Rensselaer Polytechnic Institute, Troy, NY, USA. He is currently an Associate Professor at the Electrical Engineering Program at COPPE/UFRJ, Rio de Janeiro, Brazil.

Shaopeng Wang received his B.S. in computer and system science from Nankai University, Tianjin, China in 1990, and his M.S. in electrical engineering from Beijing Univ. of Aero. and Astro., Beijing, China in 1993. He is working toward his Ph.D. degree in electrical engineering at Rensselaer Polytechnic Institute. His interests include robust control, computer aided system design.

Nelson Martins (F98) received his B.Sc. (1972) elec.eng. from Univ. Of Brasilia, Brazil, M.Sc. (1974) and Ph.D. (1978) degrees from UMIST, UK. Dr. Martins works in CEPEL, since 1978, in the development of computer tools for power system dynamics and control. He is chairman of CIGRE TF 38.02.16 on the impact of the interaction among power system controls and has contributed to several IEEE Working Groups.

Vous aimerez peut-être aussi