ABSTRACT Title of dissertation: PROBLEMS IN SELECTIVE MODAL ANALYSIS AND CONTROL Mohamed S. Saad, Doctor of Philosophy, 2007 Dissertation directed by: Professor Eyad H. Abed Department of Electrical and Computer Engineering In this dissertation, we develop monitoring and control systems for improv- ing the performance of systems that are required to operate at the edge of their stability envelopes. The concept of modal participation factors, which is an essen- tial construct in the theory of Selective Modal Analysis, is used extensively in this work. The basic deflnition of modal participation factors that was originally given for unforced linear time-invariant systems is revisited, and related notions of out- put participation factors and input-to-output participation factors are introduced, studied and applied to models of electrical power systems. A signal-based approach for real-time detection of impending instability in nonlinear systems is considered. The main idea pursued involves using a small ad- ditive white Gaussian noise as a probe signal and monitoring the spectral density of one or more measured states or outputs for certain signatures of impending in- stability. Input-to-state and input-to-output participation factors are used as tools to aid in selection of locations for probe inputs and states or outputs to be mon- itored, respectively. Since these participation factors are model-based, the work presented combines signal-based and model-based ideas toward achieving a robust methodology for instability monitoring. Case studies from power systems are used to illustrate the developed monitoring system, one of which involves the WSCC 3-generator, 9-bus network. Feedback algorithms are developed for assigning modal participation factors in general linear time-invariant systems using eigenvector assignment-based tech- niques. The goal is to reduce the interaction between a selected group of states (the high-value group) and an undesirable mode (for example a critical mode, i.e., one corresponding to an eigenvalue or pair of eigenvalues approaching the imaginary axis in the complex plane). In particular, we address two cases, one in which the mode of interest corresponds to a real eigenvalue approaching zero, and the other in which the mode involves a complex conjugate pair of eigenvalues that may be approaching the imaginary axis. A novel procedure for computing the desired closed-loop right eigenvector(s) associated with the critical mode (based on given constraints on the desired closed-loop participation factors) is presented. An example from power sys- tems is presented to demonstrate the efiectiveness of the controller. The example used is the WSCC 3-generator, 9-bus network. PROBLEMS IN SELECTIVE MODAL ANALYSIS AND CONTROL by Mohamed S. Saad Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulflllment of the requirements for the degree of Doctor of Philosophy 2007 Advisory Committee: Professor Eyad H. Abed, Chair/Advisor Professor William S. Levine Professor Gilmer L. Blankenship Professor Amr Baz Assistant Professor Richard J. La c Copyright by MOHAMED S. SAAD 2007 DEDICATION To my parents, brothers To my wife Maha and my kids Karim and Malak ii ACKNOWLEDGMENTS First and foremost I?d like to thank my advisor, Professor Eyad H. Abed for his endless support and advise throughout my graduate studies at the University of Maryland. He has been a great source of inspiration throughout the years. I would also like to thank Professors William Levine, Gilmer Blankenship, Amr Baz and Richard La for agreeing to serve on my thesis committee and for sparing their invaluable time reviewing the manuscript. I would like to thank Professor Andre Tits for the helpful discussions on my research. I would also like to thank Professor Robert Israel of the University of British Columbia for the help with my research questions. Thanks are due to Professor David Elliott for his advice on research. Thanks are also due to Dr. Monther Hasouneh for the great time we spent on discussing research and other general life topics. I owe my deepest thanks to my family - my mother and father who have always stood by me, gave me endless patience and love and guided me through my career. Words cannot express the gratitude I owe them. Deep thanks are also due to my dear brothers Ahmed and Maged. Special and deep thanks to my wife Maha. She has always been surrounding iii me with care and love. I will never forget her nice encouraging words throughout the time of my PhD. Also, I would like to thank my lovely twins, Karim and Malak for the joy they have been fllling my life with. I am grateful to all my colleagues at the University of Maryland; Chao Wu, Amirali Sharifl, Maben Rabi, Nassir BenAmmar and Guojian Lin. I will never forget our nice discussions on every aspect of life. I am also grateful to my close friends during this whole PhD journey; Mo- hamedAbdallah, MohamedZahran, Abdel-hameedBadawy, AhmedElsaid, Mostafa Arafa, Tarek Massoud and Mohamed Fahmi. I hope that our friendship will remain forever. It is impossible to remember all, and I apologize to those I?ve inadvertently left out. iv Contents List of Tables ix List of Figures xi 1 Introduction 1 1.1 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2 Mathematical Background 12 2.1 Modal Participation Factors . . . . . . . . . . . . . . . . . . . . . . . 12 2.1.1 Conventional participation factors . . . . . . . . . . . . . . . . 13 2.1.2 Generalized participation factors . . . . . . . . . . . . . . . . 16 2.2 Bifurcation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 2.2.1 Stationary bifurcation . . . . . . . . . . . . . . . . . . . . . . 19 2.2.2 Hopf bifurcation . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.3 Eigenstructure Assignment for Control System Design . . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.3.1 Eigenstructure assignment using state feedback . . . . . . . . 26 v 3 Output Participation Factors 32 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 3.2 Output Participation Factors . . . . . . . . . . . . . . . . . . . . . . 35 3.3 Input-to-Output Modal Participation Factors . . . . . . . . . . . . . 45 4 Detecting Impending Instability in Nonlinear Systems 47 4.1 Precursor-Based Monitoring . . . . . . . . . . . . . . . . . . . . . . . 48 4.1.1 Developed monitoring system . . . . . . . . . . . . . . . . . . 49 4.1.2 Instability proximity index . . . . . . . . . . . . . . . . . . . . 53 4.2 Power System Examples . . . . . . . . . . . . . . . . . . . . . . . . . 55 4.2.1 Single generator with dynamic load . . . . . . . . . . . . . . . 55 4.2.2 Single generator connected to an inflnite bus . . . . . . . . . . 57 4.2.3 Three-generator nine-bus power system . . . . . . . . . . . . . 63 5 Participation Factor Assignment Using Eigenvector Assignment- Based Feedback Control 74 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 5.2 Participation Factor Assignment: Case of Real Eigenvalue Approaching Zero . . . . . . . . . . . . . . . . . . . 77 5.2.1 Computing desired closed-loop right eigenvector ri0d . . . . . . 78 5.2.2 Feedback control design: exactly assignable case . . . . . . . . 80 5.2.3 Feedback control design: not exactly assignable case . . . . . . 84 vi 5.2.4 Feedback control design: \sacriflcial" case . . . . . . . . . . . 85 5.3 Participation Factor Assignment: Case of Eigenvalue Pair Approaching Imaginary Axis . . . . . . . . . . . . . . 90 5.3.1 Computing desired closed-loop right eigenvectors fri0d; ri+10d = ri0d?g . . . . . . . . . . . . . . . . . . . . . . . . 91 5.3.2 Feedback control design: exactly assignable case . . . . . . . . 95 5.3.3 Feedback control design: not exactly assignable case . . . . . . 98 6 Application to Electric Power Networks 101 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 6.2 Three-Generator Nine-Bus Power System . . . . . . . . . . . . . . . . 102 6.2.1 Feedback control design . . . . . . . . . . . . . . . . . . . . . 107 6.3 Remarks on Use of State Feedback in Power Networks . . . . . . . . . 112 7 Conclusions and Future Directions 114 7.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 7.2 Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 A Proof of Propositions 5.1 and 5.2 118 A.1 Proof of Proposition 5.1 . . . . . . . . . . . . . . . . . . . . . . . . . 118 A.2 Proof of Proposition 5.2 . . . . . . . . . . . . . . . . . . . . . . . . . 120 B Table of Symbols 122 vii Bibliography 123 viii List of Tables 4.1 Input-to-state participation factors and spectral peaks at !c. . . . . . 60 4.2 Parameter values for the single generator connected to an inflnite bus model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4.3 Input-to-state participation factors and spectral peaks at !c ? 7:8 for the single generator connected to an inflnite bus system. . . . . . 63 4.4 Three-generator nine-bus power system: machine data. . . . . . . . . 68 4.5 Three-generator nine-bus power system: exciter data. . . . . . . . . . 68 4.6 Input-to-state participation factors for the 3-machine nine-bus system (partial listing). The load at bus 5 is 4.4 pu. . . . . . . . . . . . . . . 68 4.7 Input-to-output participation factors for the 3-machine nine-bus sys- tem (partial listing). The load at bus 5 is 4.4 pu. . . . . . . . . . . . 69 5.1 Closed-loop ratios fz1; z2g for coe?cient vectors fk1; k2; kcadg. The desired closed-loop ratios are fzd1 = 1:2; zd2 = 2:24g. . . . . . . . . . . 100 ix 6.1 The states with the highest norms of open-loop participation factors in the critical mode associated with ?fi;i+1g = ?0:04602 ? j2:1151. The load at bus 5 is 4.4 pu. . . . . . . . . . . . . . . . . . . . . . . . 105 6.2 Closed-loop ratios fz13; z23g for coe?cient vectors fk1; k2; kcadg. The desired closed-loop ratios are fz13 = 2:92; z23 = 2:03g. . . . . . . 108 6.3 The states with the highest norms of closed-loop participation factors in the critical mode associated with ?fi;i+1g = ?0:04602 ? j2:1151. The load at bus 5 is 4.4 pu. . . . . . . . . . . . . . . . . . . . . . . . 109 x List of Figures 2.1 Transcritical bifurcation. . . . . . . . . . . . . . . . . . . . . . . . . . 21 2.2 Pitchfork bifurcation. (a) supercritical, (b) subcritical . . . . . . . . . 22 2.3 Andronov-Hopf bifurcation. (a) supercritical, (b) subcritical . . . . . 24 3.1 Illustrating diagonal transformation T in two dimensions, (a) Elliptic original uncertainty set ?, (b) Rectangular original uncertainty set ?. 43 4.1 Precursor-based instability monitor with external probe signal. . . . . 49 4.2 Power spectral densities of the states of the model given in (4.5)-(4.8). The bifurcation parameter was set to Q1 = 2:9. Band-limited AWGN of zero mean and (0:001)2 power was added to Pm. . . . . . . . . . . 58 4.3 Variation of the peak value of the power spectral density of ! as a function of the bifurcation parameter Q1. Band-limited AWGN of zero mean and (0:001)2 power was added to Pm. . . . . . . . . . . . . 59 xi 4.4 Power spectral densities of the states of the single generator connected to an inflnite bus system. The bifurcation parameter was set to KA = 185. Band-limited AWGN of zero mean and (0:000032)2 power was added to Pm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 4.5 Power spectral densities of the states of the single generator connected to an inflnite bus system. The bifurcation parameter was set to KA = 185. Band-limited AWGN of zero mean and (0:000032)2 power was added to Vref. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 4.6 Variation of the peak value of the power spectral density of VR as a function of the bifurcation parameter KA. Band-limited AWGN of zero mean and (0:000032)2 power was added to Pm. . . . . . . . . . . 66 4.7 Plot of the inverse of the power spectral density index of ! versus the bifurcation parameter KA. The threshold value is equal to 5. . . . . . 67 4.8 WSCC 3-machine, 9-bus system. . . . . . . . . . . . . . . . . . . . . . 69 4.9 Power spectral densities of the states Efd1, Efd2 and Efd3. The load on bus 5 was used as a bifurcation parameter. The load value is 4.0 pu. Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. . . . . . . . . . . . 70 xii 4.10 Power spectral densities of the states Efd1, Efd2 and Efd3. The load on bus 5 was used as a bifurcation parameter. The load value is 4.4 pu. Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. . . . . . . . . . . . 71 4.11 Power spectral density of Efd1 for three values of PL5 (the load on bus 5): PL5 = 4:0 pu (dash-dotted line), PL5 = 4:25 pu (dashed line) and PL5 = 4:4 pu (solid line). Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. 72 4.12 Power spectral densities of the state Efd1 and outputs Q14 and Q27. The load on bus 5 was used as a bifurcation parameter. The load value is 4.4 pu. Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. . . . 73 6.1 Exciter for the three-generator, nine-bus system. . . . . . . . . . . . . 104 6.2 Open-loop response for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) of the WSCC 3-gen., 9-bus power system. (a) e0q3, (b) e0q2 and (c) e0q1. . . . . . . . . . . . . 107 6.3 Plot of q-axis transient voltage of generator 1 (e0q1) versus time for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) before and after control. . . . . . . . . . . 110 xiii 6.4 Plot of q-axis transient voltage of generator 2 (e0q2) versus time for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) before and after control. . . . . . . . . . . 111 6.5 Plot of q-axis transient voltage of generator 3 (e0q3) versus time for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) before and after control. . . . . . . . . . . 112 xiv Chapter 1 Introduction This dissertation focuses on development of monitoring and feedback control sys- tems for managing the relative impact of certain critical modes on the states of a dynamical system. The work begins with a review of modal participation factors, which are quantities measuring the relative impact of a system mode in system states. Participation factors provide a convenient tool for determining which states should be monitored in order to detect that a critical mode is approaching the imag- inary axis, which would lead to instability of the system. They can also be used to determine which inputs should be used in state feedback control to mitigate the efiect of a critical mode on system states that may correspond to equipment that must be protected. We flnd that the conventional notion of participation factors is insu?cient for monitoring and control when we have access not to system states, but to system output variables. This motivates us to introduce several new con- cepts extending the traditional notion of participation factors. These include output 1 participation factors and input-to-output participation factors. We use these new concepts to develop a signal-based approach for real-time detection of impending instability in nonlinear systems. The main idea pursued involves using a small ad- ditive white Gaussian noise as a probe signal and monitoring the spectral density of one or more measured states or outputs for certain signatures of impending insta- bility. Input-to-state and input-to-output participation factors are used as tools to aid in selection of locations for probe inputs and states or outputs to be monitored. We also apply eigenstructure assignment techniques to develop procedures for as- signment of participation factors. We provide examples from the fleld of electric power systems demonstrating the concepts introduced in this thesis, both for modal monitoring and control. The conventional deflnition of participation factors was introduced in the framework of Selective Modal Analysis (SMA) in [2, 34]. SMA is a methodol- ogy for the modal analysis of linear time-invariant systems that emphasizes scale- independent notions that facilitate modal analysis, control design, and order reduc- tion. The number of modes involved in a phenomenon of interest is usually only a small fraction of those appearing in the dynamics of a system, usually conflned to a single mode. Participation factors are non-dimensional scalars that measure the interaction between modes and the state variables of a linear system. Since their introduction, participation factors have been used for stability analysis, order reduction, actuator placement and controller design in a variety of flelds, especially 2 electric power systems [9, 10, 11]. Participation factors have recently also been used in coherency and clustering studies in electrical power systems [31, 32]. In [15], a new approach to deflning modal participation factors for unforced systems was presented. This approach involves taking an average or a probabilistic expectation of a quantitative measure of relative modal participation. The average is performed with respect to all possible values of the initial state vector, assumed to be unknown but to lie in a known set or satisfy a known probability density. The deflnitions found in [15] were shown to reduce to the original deflnition of participation factors of [2, 34] if the uncertainty in the initial condition satisfles a symmetry condition. There is a need to extend the participation factors deflnitions referred to above in several directions. Two extensions that are pursued in this thesis are output participation factors and input-to-output participation factors. The need for these extensions becomes apparent when we consider systems for which the states may not be readily available for measurement or as feedback signals, while other variables (namely certain system output variables) are readily available. In such cases, it is useful to have a criterion for determining which output signal carries the most information about a critical mode of interest. In some applications, for example, clustering in power system studies, participation of modes in variables that are not state variables (e.g., bus voltages) are needed. The concept of participation of modes in output variables has not previously 3 been considered in the literature. Such a concept can be useful in many applica- tions such as clustering studies and in stability monitoring and detection of closeness to instability. In clustering in power systems, for example, development of output participation factors will facilitate studies and understanding of intracluster oscil- lations (i.e., oscillations resulting from modes within the cluster) and intercluster oscillations (i.e., oscillations resulting from modes common to more than one clus- ter [31, 32]. It will be also useful for local and coordinated control design. For example, in control design to dampen intercluster oscillations, coordinated control (distributed control) must be used. While local controls guarantee stable operation within a cluster, poor choices of local controls might lead to instabilities caused by the coupled operations of the clusters. Recently, Canizares and co-workers have proposed some extensions to the participation factors notions available in the literature, also targeting the incor- poration of output variables in the deflnitions. In [35], a concept of \extended eigenvector" was introduced and used to identify the dominant output variable as- sociated with a critical mode. In [36] an observability index was used to determine the preferred output signal for observing a critical mode. They have used their new deflnitions in a modal-based analysis technique for monitoring and control of electric power systems [35, 36]. The concept of \mode observability factor," introduced by [26], has been used to determine the best output for either monitoring a mode, or possibly as a 4 good candidate feedback signal to control a mode of interest. In particular, the mode observability factor was used for determining the most suitable buses in a power system for placing static VAR compensators in order to dampen the critical modes of oscillation. A drawback to using the deflnitions referred to above is that it is scale- dependent, i.e., for a change in the units of system outputs, the factors values will change. Another drawback to using these deflnitions is that they do not reduce to the conventional state participation factors when the output is simply a state variable. The thesis then focuses on using the newly introduced concept of input-to- output participation factors and the concept of input-to-state participation factors given in [40] to develop a signal-based approach for real-time detection of impend- ing instability in nonlinear systems with examples from power systems. Kim and Abed [25] developed monitoring systems for detecting impending instability in non- linear systems that builds on Wiesenfeld?s research on \noisy precursors of bifurca- tions". Wiesenfeld?s research was originally introduced to characterize and employ the noise ampliflcation properties of nonlinear systems near bifurcations of various types [37, 38]. Noisy precursors are features of the power spectral density (PSD) of a measured output of a system nearing instability that is excited by additive white Gaussian noise (AWGN). In [25], the noisy precursors concept was extended from systems operating at limit cycles to systems operating near equilibria, and 5 closed-loop monitoring systems were developed to facilitate use of noisy precur- sors in revealing impending loss of stability for such systems. It was shown in [25] that systems driven by white Gaussian noise and operating near an equilibrium point exhibit sharply growing peaks near certain frequencies as the system nears a bifurcation. In particular, it was shown that for stationary bifurcation where an eigenvalue passes through the origin (as in the case of pitchfork or transcritical bi- furcation), the peak in the PSD occurs at zero frequency. Analogously, for the case of Hopf bifurcation (complex conjugate pair of eigenvalues crossing the imaginary axis transversely), the peak in the PSD occurs near !c, the critical frequency of the Hopf bifurcation. As noted by Hauer [19], the recurring problems of system oscillations and voltage collapse in power systems are due in part to system behavior not well cap- tured by the models used in planning and operation studies. In the face of compo- nent failures, system models quickly become mismatched to the physical network, and are only accurate if they are updated using a powerful and accurate failure detection system. Therefore, it is important to employ nonparametric techniques for instability monitoring. The use of probe signals is shown to help reveal an impending loss of stabil- ity. This is because probe signals propagate in the power system and give certain signatures near an instability that can be used as a warning signals for possible impending voltage collapse. Such warning signals are needed to alert system opera- 6 tors of a situation that may require preventive control, and to provide the operators with valuable additional time to take necessary preventive (rather than corrective) measures. The focus of the thesis then shifts to addressing possible actions to be taken to counter the possible efiects of the impending instability on the system. The need to carry such actions can be attributed to the fact that in many systems composed of smaller subsystems, it may be desirable to assign the magnitudes of participation factors corresponding to various subsystems to desired values. This would ensure that the energy corresponding to a critical mode minimally afiects subsystems of high value, directing the energy of this mode to other, possibly purposefully intro- duced, subsystems of lesser value to the system operator. This idea is based on the fact that the higher the magnitude of participation factor, the higher the tendency of a subsystem to be afiected in a serious way by an impending (or true) instability. Power systems are a good example for this discussion. In power systems operation, nuclear units, for example, are regarded as high-value subsystems. Reducing partic- ipation factors of these units in some critical mechanical mode would reduce the risk of premature tripping, as well as alleviate the efiect on these units of a troublesome mode being excited. The goal of the control efiort will not be to alter the position of the eigen- value(s) (associated with the critical mode), but rather to reduce the participation factors of a high-value group of states in that mode. This means that the closed-loop 7 eigenvalues will be kept the same as the open-loop ones. Similarly, all open-loop right eigenvectors will be kept unchanged except for those eigenvector(s) associated with the critical mode. This is possible using the freedom ofiered by state feedback in multivariable systems beyond closed-loop eigenvalue assignment [17]. Moreover, and as noted in the seminal paper of Moore [17], \One advantage of eigenstructure assignment is that you can still assign eigenvectors associated with uncontrollable eigenvalues." This is important in participation factor assignment where it is desired that open-loop eigenvalues are kept unchanged. A state feedback control system is given for assigning modal participation factors using eigenvector assignment-based techniques. The idea of using eigenvector assignment-based techniques to assign participation factors has not previously been addressed. Some work has addressed the problem of assigning right eigenvector(s) associated with a mode of interest (usually the critical mode). In [4], the author alerted the power systems community to the possible use of eigenvector assignment to make a target area participate highly in an unstable mode. This could be done by an unscrupulous provider of power to hurt the operation of a competing provider. In [13], a control scheme was proposed to assign eigenvalues and eigenvectors to improve system damping and dynamic behavior, respectively. Since right and left eigenstructure problems are strongly interdependent, assigning participation factors is not an easy task. Indeed, if the required objective is not exactly attainable, participation factor assignment may require solving an optimization problem for 8 achieving an objective as close as possible to the desired one. Inspired by vibration absorbers in mechanical and civil engineering, in this work we desire to design controllers that reduce participation of certain, \high- value" system states in a particular mode at the expense of other \low-value" states. Those low-value states can be viewed as taking a role akin to that of vibration absorbers in mechanical systems. In power systems for instance, an isolated remote generator station or ywheel could be regarded as a vibration absorber that protects heavilyloadedhigh-capacitygeneratingstations. Wewillrefertotheaforementioned problem as the participation factor assignment problem. Reducing the participation factors of a high-value group of states in the critical mode implies that some other groups of states will be burdened with higher magnitudes of their participation factors in the critical mode. Deciding which group of states will be seen as \low-value" is based on system conditions and operator experience. Control systems that assign participation factors should work in unison with monitoring systems that detect closeness to instability. The control becomes active when the monitoring system provides an alert that instability is near. The objective of participation factor assignment in this setting is as described above: to lessen the impact of a critical mode on high-value states by a trade with impact on low-value states. 9 1.1 Thesis Outline The dissertation proceeds as follows. In Chapter 2, theoretical background mate- rial used throughout the thesis is collected. The topics include bifurcation theory, modal participation factors and eigenstructure assignment-based feedback control algorithms. In Chapter 3, a study is given on extending the uncertain initial condition approach to deflning state participation factors in [15] to facilitate flnding a general measure of output participation. The new proposed deflnition has the desirable property that it reduces to the original state participation deflnition when the out- puts are simply the system states. Also, this new deflnition is unit independent, i.e., the output participation factors do not depend on the units of system states or outputs. Another notion of modal participation is presented in this chapter. This new deflnition is one of input-output participation factors. It involves probe input signals as well as modes and outputs, and is applicable to a large set of systems. In Chapter 4, a signal-based approach for real-time detection of impending instability in nonlinear systems is considered. The developed monitoring system is based on using a small additive white Gaussian noise as a probe signal and monitor- ing the spectral density of one or more measured states or outputs for certain sig- natures of impending instability. Input-to-output participation factors (introduced in Chapter 3), and input-to-state participation factors [40] (discussed in Chapter 2) are used as tools to aid in selection of locations for probe inputs and outputs or 10 states to be monitored. Examples from power systems are given to demonstrate the performance of the monitoring system. In Chapter 5, state feedback control is presented for assigning modal partic- ipation factors for systems nearing instability using eigenvector assignment-based techniques. The focus is on assigning participation factors for state variables within a predeflned group of state variables, where the group can be any subvector of the state vector, or could consist of all system states. A procedure for computing the desired closed-loop right eigenvector(s) (based on a given desired closed-loop par- ticipation factors) is given. Two cases are addressed, flrst the case of systems with a mode associated with a single real eigenvalue approaching zero, and second the case of systems with a mode associated with a pair of conjugate complex eigenvalues approaching the imaginary axis. In Chapter 6, the results obtained in Chapter 5 are applied to a sample power system operating close to its stability limits. The considered power system is the WSCC 3-generator, 9-bus network. As expected, participation factor assignment has a clear efiect on the relative time domain responses of system states. Finally, we collect concluding remarks in Chapter 7, along with possible direc- tions for future research. Some of the results reported in this thesis were published in a conference paper and a book chapter [22, 21]. 11 Chapter 2 Mathematical Background In this chapter, the theoretical and algorithmic background materials employed throughout this dissertation are reviewed. Section 2.1 focuses on modal partici- pation factors, a basic construct in the theory of Selective Modal Analysis [2, 34]. In Section 2.2 a summary on bifurcation theory is provided, and various types of continuous-time local bifurcations are reviewed. In section 2.3, background material on eigenstructure assignment for control system design is reviewed. 2.1 Modal Participation Factors In the theory of Selective Modal Analysis [2, 34], participation factors are non- dimensional scalars that measure the interaction between the modes and the state variables of a linear time-invariant system. The notion of participation factors pro- vides a tool for modal participation analysis where an e?cient method is required to single out a selected portion of the model related to the dynamics of interest. The 12 conventional deflnition of participation factors has been used as a tool for stability analysis, modal reduction, actuator placement and controller design in a variety of flelds, especially electric power systems [9, 11, 10]. The number of modes involved in the phenomenon of interest is usually only a small fraction of those governing the dynamics of the system, mostly restricted to only one mode. Participation factors have also recently been used in coherency and clustering studies in electrical power systems [31, 32]. Another potential use of modal participation factors is in the area of stability monitoring of nonlinear systems close to instability. Such systems are also known to be close to a bifurcation in their dynamics. Input-to-state participation factors (given in [40]) and input-to-output participation factors (introduced in Chapter 3) will be employed in Chapter 4 to design stability monitoring systems for detecting impending instability in nonlinear systems. 2.1.1 Conventional participation factors Consider a general continuous linear time-invariant system _x(t) = Ax(t) (2.1) where x(t) 2 Rn, and A is a real (n ? n) matrix. Suppose that A has a set of n distinct eigenvalues f?1;?2;:::;?ng. Let fr1;r2;:::;rng be right eigenvectors of the matrix A associated with the eigenvalues f?1;?2;:::;?ng, respectively. Let fl1;l2;:::;lng denote left (row) eigenvectors of the matrix A associated with the 13 eigenvalues f?1;?2;:::;?ng, respectively. The right and left eigenvectors are taken to satisfy the normalization lirj = ?ij (2.2) where ?ij is the Kronecker delta: ?ij = 8 >>< >>: 1 i = j 0 i 6= j The participation factor of the i-th mode in the k-th state is deflned to be the complex number [2, 34, 15] pki := likrik (2.3) This formula also gives the participation of the k-th state in the i-th mode. Partic- ipation factors is a measure of the relative participation of modes in states and the relative participation of states in modes. Since rik measures the activity of xk in the i-th mode and lik weighs the contribution of this activity to the mode, the product pki measures the net participation. Properties of participation factors The following properties of participation factors are recalled from [1, 3] and can be easily proved through appropriate consideration of the Deflnition in (2.3) and the normalization assumption in (2.2). 1. Participation factors are dimensionless quantities that are independent of the units in which state variables are measured. 14 2. In view of the eigenvector normalization (2.2), the sum of the participation fac- tors associated with any mode (Pni=1 pki) or with any state variable (Pnk=1 pki) is equal to 1. 3. The participation factor pki represents the sensitivity of the eigenvalue ?i to the the diagonal element akk of the state matrix A pki = @?i@a kk : The solution to system dynamic equations in (2.1) for an initial condition x0 is given by: x(t) = eAtx0 (2.4) Since the state matrix A is assumed to have distinct eigenvalues, then A could be written in terms of right and left modal matrices as well as a diagonal matrix of its eigenvalues. Then x(t) in (2.4) is x(t) = nX i=1 (lix0)e?itri (2.5) Now suppose the initial condition x0 is ek, the unit vector along the k-th coordinate axis. Then the evolution of the k-th state becomes xk(t) = nX i=1 (likrik)e?it (2.6) = nX i=1 pkie?it: (2.7) 15 Equation (2.7) indicates that pki can be viewed as the relative participation of the i-th mode in the k-th state at time t = 0 1. 2.1.2 Generalized participation factors The concept of participation factors of modes in states and vice versa has been extended to linear time-invariant systems with inputs [40] _x(t) = Ax(t)+Bu(t) (2.8) y(t) = Cx(t): (2.9) In [40], the case where the input is applied to one component, say the q-th compo- nent, of the right side of (2.8) and only one state, say the k-th state, is measured is considered. That is, in equations (2.8)-(2.9), B and C take the form B = eq = [0 ::: 0 1|{z} q?th 0:::0]T; C = ekT = [0 ::: 0 1|{z} k?th 0:::0]: 1In [15], a new approach to deflning modal participation factors was presented. The new approach involved taking an average or a probabilistic expectation of a quantitative measure of relative modal participation over an uncertain initial state vector. The new deflnitions were shown to reduce to the original deflnition of participation factors of [2, 34] if the initial state obeys a symmetry condition. 16 With this choice of C and B, in steady state the output in (2.9) (in the frequency domain) is given by y(s) = xk(s) = C(sI ?A)?1Bu = nX i=1 CriliB s??i u(s) = nX i=1 rikliq s??iu(s) The generalized participation factor, participation factor of mode i in state k when excitation is applied to state q, is deflned as follows: piqk := r i kl i q kHkk (2.10) Here kHkk := jri~kri~qj is the normalization factor where (~k; ~q) = arg maxp;q (jrikliqj) attains maximum participation of norm 1. The above normalization makes the participation factors to be bounded by zero and one in amplitude and independent of system scaling. In [1], the quantity piqk = rikliq is called a generalized participation. We will call this quantity the input-to-state participation factor (ISPF) for mode i, with measurement at state k and input applied to state q. The input-to-state participation factor (ISPF) will be employed in designing monitoring systems of impending instabilities in nonlinear systems in Chapter 4. 17 2.2 Bifurcation Theory The term bifurcation is commonly used to describe a qualitative change in steady state behavior as the parameters on which the dynamical system depends are var- ied through certain values [6, 7, 5]. The parameter being varied is referred to as the bifurcation parameter. In other words, a bifurcation is a change in the num- ber of candidate operating conditions of a nonlinear dynamical system that occurs as a parameter is quasi-statically varied. A value of the bifurcation parameter at which a bifurcation occurs is called a critical value of the bifurcation parameter. A nonlinear system operating at an equilibrium point undergoes a bifurcation when a quasistatic change in parameters causes the equilibrium to lose stability. Typically, new equilibria and/or limit cycles appear from the nominal equilibrium at a bifur- cation. When these so-called bifurcated solutions are stable, the bifurcation is said to be supercritical or soft. When they are unstable, the bifurcation is said to be subcritical or hard. The implication of these concepts for operation of real systems is easy to describe (see [5] for a detailed exposition). For a subcritical bifurcation, the system trajectory escapes from a local neighborhood of the nominal equilibrium and can diverge to a distant (unacceptable) steady state. Therefore, detecting an impending subcritical bifurcation is highly desirable. For a supercritical bifurcation, on the other hand, the motion is likely to settle on the nearby bifurcated steady state, which is stable and does not represent a major departure from the nominal operation. 18 To flx ideas, consider a general one-parameter family of ordinary difierential equation systems _x = F?(x;?) (2.11) where x 2Rn is the system state, ? 2Rdenotes the bifurcation parameter, and F? is smooth in x and ?. For any value of ?, the equilibrium points of (2.11) are given by the solutions for x of the algebraic equations F?(x;?) = 0. Local bifurcations are those that occur in the vicinity of an equilibrium point. Local bifurcations of equi- librium points consist of stationary bifurcations and the Andronov-Hopf bifurcation (Hopf bifurcation for short). Stationary bifurcation is any bifurcation of one or more equilibrium points from a nominal equilibrium point. Saddle-node, transcriti- cal, and pitchfork bifurcations are some examples of stationary bifurcation. Unlike the stationary bifurcation, in Hopf bifurcation a branch of periodic orbits bifurcates from an equilibrium point. Stationary and Hopf bifurcation of equilibrium points can be also categorized on the basis of the critical eigenvalues at the corresponding bifurcation points. Next, a brief summary of various types of continuous-time local bifurcations is given. 2.2.1 Stationary bifurcation Stationary bifurcation occurs when a single real eigenvalue goes from being negative to being positive as the bifurcation parameter ? passes through a critical value ?c. Precisely, the origin of (2.11) undergoes a stationary bifurcation at the critical 19 parameter value ? = 0 if assumptions, (L1), (L2) and (L3) hold. (L1) F? of system (2.11) is su?ciently smooth in x, ?, and F?(0) = 0 for all ? in a neighorhood of 0 (i.e., 0 is an equilbrium for all values of ? in the neighborhood); (L2) The Jacobian A(?) := DxF?(0) possesses a simple real eigenvalue ?(?) such that at the critical parameter value ?c = 0, ?(0) = 0 and ?0(0) := @?@?j?=0 6= 0; (L3) All eigenvalues of the critical Jacobian A(0) besides 0 have negative real parts. Under assumptions (L1), (L2) and (L3), two new equilibrium points of (2.11) emerge from the origin at ? = 0. Bifurcation stability coe?cients are quantities that determine the direction of bifurcation, and in particular the stability of the bifurcated solutions. Locally, the new equilibrium points occur for parameter values given by a smooth function of an auxiliary small parameter ? (? can be positive or negative): ?(?) = ?1?+?2?2 +O(?3) (2.12) One of the new equilibrium points occurs for ? > 0 and the other for ? < 0. Also, the stability of the new equilibrium points is determined by the sign of an eigenvalue fl(?) of the system linearization at the new equilibrium points. This eigenvalue is near 0 and is also given by a smooth function of the parameter ?: fl(?) = fl1?+fl2?2 +O(?3) (2.13) 20 Stability of the bifurcated equilibrium points is determined by the sign of fl(?). If fl(?) < 0 the corresponding equilibrium point is stable, while if fl(?) > 0 the equilibrium point is unstable. The coe?cients fli, i = 1;2;::: in the expansion above are the bifurcation stability coe?cients mentioned earlier, for the case of stationary bifurcation. The values of these coe?cients determine the local nature of the bifurcation, as explained next. Since ? can be positive or negative, it follows that if fl1 6= 0 the bifurcation is neither subcritical nor supercritical. (This is equivalent to the condition ?1 6= 0.) The bifurcation is therefore generically transcritical. Figure 2.1 shows the corre- sponding bifurcation diagram. If fl1 = 0 and fl2 6= 0, a stationary bifurcation is known as a pitchfork bifurcation. The pitchfork bifurcation is subcritical if fl2 > 0; it is supercritical if fl2 < 0. Figure 2.2 shows the corresponding bifurcation diagram. x ? Figure 2.1: Transcritical bifurcation. 21 x ? (b) (a) x ? Figure 2.2: Pitchfork bifurcation. (a) supercritical, (b) subcritical 2.2.2 Hopf bifurcation Suppose that the origin of (2.11) loses stability as the result of a complex conjugate pair of eigenvalues of A(?) crossing the imaginary axis. All other eigenvalues are assumed to remain stable, i.e., their real parts are negative for all values of ?. Under this simple condition on the linearization of a nonlinear system, the nonlinear system typically undergoes a bifurcation. To be more precise, assume the following conditions are satisfled (the critical parameter value is taken to be ?c = 0 without loss of generality). (H1) F? of system (2.11) is su?ciently smooth in x, ?, and F?(0;?) = 0 for all ? in a neighborhood of 0. The Jacobian A(?) possesses a complex-conjugate pair of (algebraically) simple eigenvalues ?1;2(?) = fi(?) ? i!(?), such that fi(0) = 0, 22 fi0(0) 6= 0 and !c := !(0) > 0. (H2) All eigenvalues of the critical Jacobian A(0) besides ?i!c have negative real parts. The Hopf Bifurcation theorem asserts that, under conditions (H1) and (H2), a small- amplitude non constant limit cycle (i.e., periodic solution) of (2.11) emerges from the origin at ? = 0. Locally, the limit cycles occur for parameter values given by a smooth and even function of the amplitude ? of the limit cycles: ?(?) = ?2?2 +?4?4 +::: (2.14) where the ellipsis denotes higher order terms. The stability of the limit cycle result- ing from a Hopf bifurcation is determined by the sign of a particular characteristic exponent fl(?). This characteristic exponent is real and vanishes in the limit as the bifurcation point is approached. It is given by a smooth and even function of the amplitude ? of the limit cycles: fl(?) = fl2?2 +fl4?4 +::: (2.15) The coe?cients ?2 and fl2 in the expansions above are related by the exchange of stability formula fl2 = ?2fi0(0)?2: Generically, these coe?cients do not vanish. Their signs determine the direction of bifurcation. The coe?cients fl2, fl4;::: in the expansion (2.15) are the bifurcation 23 stability coe?cients for the case of Hopf bifurcation. If fl2 > 0, then locally the bifurcated limit cycle is unstable and the bifurcation is subcritical. If fl2 < 0, then locally the bifurcated limit cycle is stable (more precisely, one says that it is orbitally asymptotically stable). This is the case of supercritical Hopf bifurcation. If it happens that fl2 vanishes, then stability is determined by the flrst non vanishing bifurcation stability coe?cient (if one exists). Figure 2.3 shows a typical Hopf- bifurcation diagram. y x ? (b) (a) y x ? Figure 2.3: Andronov-Hopf bifurcation. (a) supercritical, (b) subcritical 24 2.3 Eigenstructure Assignment for Control System Design Eigenstructure assignment problem is considered the general extension to the well known eigenvalue assignment problem [23, 17]. This general notion is attributed to the fact that, apart from the case of single-input single-output (SISO) systems, the speciflcation of closed-loop eigenvalues does not uniquely deflne the feedback structure of a closed-loop system. The source of nonuniqueness can be identifled as that coming from the freedom ofiered by state or output feedback in multi-input multi-output (MIMO) systems, beyond eigenvalue assignment, in selecting the as- sociated eigenvectors and generalized eigenvectors from allowable spaces. Most of the work on eigenstructure assignment has been purely eigenvalue assignment (pole placement) up to late 60?s. Later, Moore found that degrees of freedom are available over and above eigenvalue assignment using state feedback control for linear time- invariant multi-input multi-output (MIMO) systems [17]. Since then, numerous methods and algorithms have been developed to exercise those extra degrees of free- dom to give the system some good performance characteristics. Good performance characteristics span the areas of low sensitivity, robust, multi-objective and fault detection [23]. In this section, background material on eigenstructure assignment that is needed in the dissertation is recalled. 25 2.3.1 Eigenstructure assignment using state feedback Consider the closed-loop LTI system _x(t) = (A+BF)x(t) (2.16) associated with an open-loop system _x(t) = Ax(t)+Bu(t) (2.17) to which a state feedback u(t) = Fx(t) is applied. Here x(t) 2 Rn; u(t) 2 Rm; A 2R(n?n); F 2R(m?n). For a given complex scalar ?, which may be viewed as a closed-loop eigenvalue, deflne the Hautus matrix [4]: S? := ? (?I ?A) B ? ; S? 2Cn?(n+m) Also deflne the matrix K?, with its row partitions along the same boundary as the partitions of columns in S?. K? is given by: K? := 2 66 4 N? M? 3 77 5; K? 2C(n+m)?m; (2.18) and is composed of columns that constitute a basis for null space (or kernel) of S?. The matrix K? can be calculated by performing the singular value decomposition (SVD) or orthogonal triangular decomposition (OTD) on S? [23]. To compute K? through SVD, apply singular value decomposition to S? to yield: S? = U?V H; (2.19) 26 where U is an (n?n) unitary matrix over the fleld of complex numbers, the matrix ? is (n?(n+m)) with nonnegative numbers (singular values of S?) on the diagonal and zeros ofi the diagonal, and V H denotes the conjugate transpose of V, an ((n+ m)?(n+m)) unitary matrix over the fleld of complex numbers. The matrix K? is given by the last m columns of the unitary matrix V. Knowing K? means that for any z 2C(m+n) such that S?z = 0; (2.20) there exists a k 2Cm such that z = K? k: (2.21) It is clear that the columns of N? are linearly independent if B has linearly independent columns, and that N?? = N??. It is known that F may be chosen to yield any self-conjugate set of closed-loop eigenvalues provided that the pair (A;B) is controllable, i.e., rank [B AB ::: An?1B] = n Moore in [17] described the available freedom for assigning eigenvectors for an arbitrary self-conjugate set of eigenvalues using state feedback in the case m > 1. He gave both necessary and su?cient conditions for a state feedback to exist that achieves a desired set of eigenvalues and eigenvectors. Proposition 2.1 [17] Consider f ?1;?2;:::;?n g a set of self-conjugate distinct com- plex numbers, and a set of complex vectors fv1;v2;::::;vng. There exists a matrix F 27 of real numbers such that ?ivi = (A + BF)vi for all i 2 f1;2;:::;ng if and only if the following three conditions are satisfled: 1. The vectors fv1;v2;:::;vng form a linearly independent set within Cn (over the fleld of complex scalars), 2. vi = vj? whenever ?i = ??j, and 3. vi 2 span fN?ig, where N?i is deflned as given in (2.18). Furthermore, if such a feedback gain matrix F exists and rank B = m, then F is unique. A proof of this proposition, along with a procedure for computing the feed- backgain matrix F, is givenin [17]. Next, werecall Moore?s procedure for computing the feedback gain matrix F [17]. Computing feedback gain matrix F Suppose that vi, i 2f1;2;:::;ng are chosen to satisfy the three conditions of Propo- sition 2.1. Since vi 2 span fN?ig for i 2 f1;2;:::;ng, then vi can be expressed as vi = N?i ki: (2.22) 28 for some vector ki 2Rm(Cm) (based on whether ?i is real (complex)), which implies that ? (?I ?A) B ? 2 66 4 vi M?iki 3 77 5 = 0; i.e., (?iI ?A)vi +BM?iki = 0: If F is chosen such that ?M?iki = Fvi, then [?iI ?(A+BF)]vi = 0. Let wi = ?M?i ki: If all n eigenvalues are real numbers, then vi; wi are vectors of real numbers. Also the flrst condition of Proposition 2.1 implies that the matrix [v1 v2 :::vn] is always nonsingular. For this case, F = ? w1 w2 ::: wn ? ? v1 v2 ::: vn ??1 (2.23) For the case where there are complex eigenvalues, assume that ?1 = ??2. The second condition of Proposition 2.1 states that v1 = v2? which implies that w1 = w2?. The equation which must be solved is then: F ? v1R +jv1I v1R ?jv1I V ? = ? w1R +jw1I w1R ?jw1I W ? (2.24) where the columns of V and W are vi, i = 3;:::;n and wi, i = 3;:::;n, respectively. Multiplying both sides of (2.24) from the right by the nonsingular matrix 29 2 66 66 4 1 2 ?j 1 2 0 1 2 +j 1 2 0 I 3 77 77 5 yields the equivalent equation F ? v1R v1I V ? = ? w1R w1I W ? : As vi, i 2 f1;2;:::;ng are independent, the columns of [v1R v1I V] are linearly independent. Then F is given by: F = ? w1R w1I W ? ? v1R v1I V ??1 : (2.25) The third condition of Proposition 2.1 implies that the eigenvector vi must be in the subspace spanned by the columns of N?i. This subspace is of dimension m which is equal to the rank of B, or equivalently the number of independent control variables. Therefore, the number of control variables available determines how large (dimension) the subspace in which achievable eigenvectors must reside. The orientation of the subspace is determined by the open-loop quantities A; B and the desired closed-loop eigenvalue ?i. In general, however, a desired eigenvector vid will not reside in the prescribed subspace and hence cannot be achieved. Instead a \best possible" choice for an achievable eigenvector is made. Next, we recall the analysis to compute the best possible achievable eigenvector via associated with a desired eigenvector vid [8]. 30 \Best possible via" Thisbestpossibleachievableeigenvectorvia istheprojectionofvid ontothesubspace spanned by the columns of N?i. To compute via associated with a real eigenvalue ?i, recall from (2.22) that via = N?i ki; ki 2Rm: To flnd the value of ki corresponding to the projection of vid onto the \achievable subspace," ki is chosen to minimize J = kvid ?viak2 = kvid ?N?i kik2: Now dJ=dki = 2 NT?i(N?i ki ?vid): Hence dJ=dki = 0 implies ki = (NT?i N?i)?1 NT?i vid (2.26) via = N?i (NT?i N?i)?1 NT?i vid: (2.27) For complex eigenvalues/eigenvectors the presentation above is formally correct with minor arithmetic adjustments to accommodate complex quantities [17]. 31 Chapter 3 Output Participation Factors In this chapter, the basic deflnition of modal participation factors that was originally given for LTI systems without input or output signals [2, 34] is revisited, along with an extension proposed in [15], and then several new extensions are proposed. The flrst such extension is a concept of output participation factors that measure the participation of modes in output variables that are not necessarily system states. The deflnition is based on computing an average relative contribution of a mode in an output relative to uncertainty in the system initial condition. This follows the approach proposed in [15] for deflning state participation factors. The new deflnition proposed here has the desirable property that it reduces to the original state participation deflnition when the outputs are simply the system states. Also, this new deflnition is unit independent, i.e., the output participation factors do not depend on the units of system states or outputs. It is also found that the proposed output participation factors depend on the set to which the uncertain 32 initial condition belongs. Another notion of modal participation is presented in this chapter. This new deflnition is one of input-output participation factors. It involves probe input signals as well as modes and outputs, and is applicable to a large set of systems. This chapter proceeds as follows. In Section 3.1, we recall the work of [15] on deflning participation factors using averaging over an uncertain initial condition. In Section 3.2, development of a notion of output participation factors is presented. In Section 3.3, input-to-output participation factors are introduced and studied. 3.1 Introduction Consider a continuous linear time-invariant system described as given in (2.1). The system solution for a general initial condition x0 (given in (2.5)), and the solution for the particular initial condition x0 = ek (given in (2.7)) are repeated here for convenience: _x(t) = Ax(t) (3.1) x(t) = nX i=1 (lix0)e?itri (3.2) x(t) = nX i=1 (likrik)e?it = nX i=1 pkie?it: (3.3) 33 Abed et al. [15] redeflned the participation factors concept of Verghese et al. [2, 34] by considering the efiect of uncertainty in the initial condition x0. The reconsid- eration of modal participation in this light was motivated by the following simple observations for the linear system (3.1) and its solution (3.2): If the initial condi- tion x0 lies along the i-th eigenvector, then the only mode that participates in the evolution of any state is the i-th mode. On the other hand, if the initial condition lies along the k-th coordinate axis, then the evolution of the k-th state involves all system modes according to (3.3). Clearly, then, the degree to which a mode par- ticipates in a state depends on the initial condition. The authors proposed a new deflnition for state participation factors that reduces to the original deflnition of participation factors of [2, 34] if the initial condition x0 is taken to lie in a connected set ? containing the origin and obeys a symmetry condition deflned as follows: Deflnition 3.1. ([15]) The set ? is symmetric with respect to each of the hy- perplanes xk = 0, k = 1;2;:::;n if and only if for any k 2 f1; :::;ng and z = (z1;:::;zk;:::;zn) 2Rn, z 2 ? implies that (z1;:::;?zk;:::;zn) 2 ?. The symmetry assumption in Deflnition 3.1 is reasonable for typical engineering sys- tem models although some applications might have restrictions on initial conditions. The deflnition of state participation factors proposed in [15] is: Deflnition 3.2. ([15]) The participation factor for the mode associated with ?i in 34 state xk with respect to an uncertainty set ? is pki := avg x02? f(l ix0)ri k x0k g (3.4) whenever this quantity exists. Here, avgx02? is an operator that computes the average of a function over the set ?. This quantity measures the average relative contribution at time t = 0 of the i-th mode to state xk . In the deflnition, the i-th mode is interpreted as the e?it term in (3.2). Also, the denominator on the right-hand side of (3.4) is simply the sum of the contributions of all system modes to xk(t) at t = 0: x0k = nX j=1 (ljx0)rjk: (3.5) In the next section, we consider the natural extension of the deflnition above to a concept of participation of modes in outputs. The efiect of uncertainty in the initial condition x0 is incorporated in a way similar to that used in [15] for calculation of state participation factors. 3.2 Output Participation Factors Consider a continuous linear time-invariant system with outputs _x(t) = Ax(t) (3.6) y(t) = Cx(t) (3.7) 35 where x 2Rn, y 2Rm, A is a real n?n system matrix and C is a real m?n output matrix. Using the solution for x(t) in (3.2), the output vector is given by y(t) = C nX i=1 (lix0)e?itri (3.8) The evolution of the k-th output variable yk is given by yk(t) = Ck nX i=1 (lix0)e?itri (3.9) where Ck is the k-th row of C. It may be tempting to say that the participation of the i-th mode in the k-th output is equal to the k-th row of C multiplied by the state participation factors, i.e., the sum of the participation of the i-th mode in all states weighted by the elements of the k-th row of C: pyik = nX j=1 Ckj pji (3.10) This temptation is supported by the fact that the expression in (3.10) reduces to the participation of state k in mode i when Ck = ek (the k-th unit vector). However, this would lead to a notion that is unit-dependent and hence unacceptable from the point of view of this dissertation. Next, we give a high-level deflnition that is the natural extension of the state participation factors deflnition in [15] to the case of participation of modes in outputs. It is easily verifled that this deflnition is independent of units. Below, the concept is reduced to a simple calculation for systems with initial condition uncertainty known to lie in an n-ellipsoid or n-rectangle with main axes parallel to the coordinate axes. 36 Deflnition 3.3. The participation factor for the mode associated with ?i in output yk with respect to an uncertainty set ? is pyik := avg x02? f(l ix0)(Ckri) y0k g (3.11) whenever this quantity exists. Here, the expression avgx02? denotes the operator that computes the average of a function over the set ?. Note that in computing the implied multidimensional volume integral, the argument function is undeflned for y0k = 0, i.e., undeflned for x0 2 fx : Ckx = 0g and the Cauchy principal value of the integral is to be used. The quantity on the right-hand side of Deflnition 3.3 measures the average relative contribution at time t = 0 of the i-th mode to output yk . The denominator of this quantity is simply the sum of the contributions of all modes to yk(t) at t = 0: y0k = nX j=1 (ljx0)Ckrj = Ckx0: Writing li = liCk + li?Ck, where liCk is the projection of li in the direction of Ck and li?Ck is the projection of li in the direction orthogonal to Ck, the expression in (3.11) becomes pyik = Ckri avg x02? f(l i Ck +l i? Ck)x 0 Ckx0 g = Ckri avg x02? fl i Ckx 0 Ckx0 + li?Ckx0 Ckx0g (3.12) 37 Writing liCk in terms of li and Ck, liCk = klik?f l iCkT klikkCkkg?f Ck kCkkg = fl iCkT kCkk2g?C k (3.13) Using (3.13), pyik reduces to pyik = Ckri avg x02? fl iCkT kCkk2 + li?Ckx0 Ckx0g = C kriliCkT kCkk2 +C kri avg x02? fl i? Ckx 0 Ckx0g Let Vol(?) := Z x02? dx0 (3.14) denote the volume of the set ?. The expression on the right-hand side of (3.11) is pyik = C kriliCkT kCkk2 +C kri Z x02? fl i? Ckx 0 Ckx0gdx 0=Vol(?) (3.15) The orthogonal vectors li?Ck and Ck operate on the initial state vector x0 2 ? to orthogonally transform it to another set ?0. Generally, even if the set ? is symmetric according to Deflnition 3.1, the new set ?0 is not symmetric with respect to the axes x0i = 0, i 2 f1;2;:::;ng, and the integral term doesn?t vanish. Note that in the case where output yk is simply the k-th state (i.e., Ck = ek), the integral term on the right hand side of (3.15) vanishes. This follows from the observation that the integral term reduces to g ? Z x02? fx 0 j x0kgdx 0 38 for some j 6= k and a constant g = Ckri=Vol(?). Because ? is symmetric according to Deflnition 3.1, Z x02? fx 0 j x0kgdx 0 = 0 where the integral is interpreted in the sense of the Cauchy principal value. The expression in (3.11) is dimensionless, i.e., it does not depend on the units in which state variables are measured or the unit in which the output variable is measured. Also, its simpliflcation (3.15) reduces to the conventional state participation factors when the output is simply a state variable. Thus, it can be used to provide a good measure of the the mode that has the highest participation in an output when it is reasonable to assume uncertainty in the system?s initial conditions. Also note that the deflnition in its current form is dependent on the set ? to which x0 belongs. Suppose that the initial condition uncertainty set ? is known to be symmetric according to Deflnition 3.1. This is not enough to ensure that the integral term in (3.15) vanishes. However, this integral vanishes if the initial condition uncertainty set is an n-cube or an n-sphere. For instance, the initial state z0 lies in an n- sphere if kz0k ? c0 where c0 is a positive constant. Therefore, suppose the initial condition uncertainty set is an n-ellipsoid or n-rectangle with main axes parallel to the coordinate axes. Next, consider a change in units (x0 = Tx), T is a positive diagonal matrix, such that in the new coordinates the state initial condition x00 lies either in an n-cube or an n-sphere. In this case, the integral term vanishes (a proof 39 is given below) and the expression in (3.15) reduces to pyik = C 0kr0il0iC0kT kC0kk2 ; (3.16) where C0k = CkT?1; r0i = Tri and l0i = liT?1: Proof: Let x00 = Qz0, where Q is an (n?n) orthonormal matrix, i.e., Q?1QT = I The orthonormal similarity transformation matrix Q can be written as Q = 2 66 4 ^C0k Q1 3 77 5 T where ^C0k = C0k kC0kk; and Q1 is chosen such that C0kQT1 = 0. Note that kx00k = kz0k, dx00 = dz0. Using the orthonormal similarity transformation Q, the integral term in (3.15) becomes Z x00?c0 fl 0i? C0k x 00 C0k x00 gdx 00 = Z z0?c0 fl 0i? C0k Q z 0 C0k Q z0 gdz 0 = Z z0?c0 fl 0i? C0k ( ^C0 k)T z0 1 +l i? C0k Q T 1 z 00 C0k ( ^C0k)T z01 gdz0 = Z z0?c0 f l 0i? C0kQ T 1 z 00 C0k ( ^C0k)T z01 gdz0 = 0; (3.17) 40 where z0 = [z01 z00]T. The last step in (3.17) follows from the observation that for any j 6= 1, Z z0?c0 fz 0 j z01gdz 0 = 0; where the integral above is interpreted in the sense of Cauchy principal value. ? Eq. (3.16) is taken as the deflnition of output participation factors, and is applied to the transformed system in which the initial condition uncertainty set is an n-cube or n-sphere. The expression in (3.16) is crucial for computing the value of the high-level deflnition of output participation factors given in Deflnition 3.3. This value (that depends on the n-ellipsoid or n-rectangle set ? with main axes parallel to the co- ordinate axes) is equal to the value of the deflnition of output participation factors in (3.16) when a diagonal transformation (equivalent to a change in units of system states) is performed to transfer the set ? into an n-cube or an n-sphere. Next, we summarize the properties of output participation factors (deflned in (3.16)). These properties can be easily proved through appropriate consideration of Deflnition 3.3 and Eq. (3.16). Properties of output participation factors 1. The value of the high-level deflnition of output participation factors given in Deflnition 3.3, and which is a function of A; C and the set ?, is equal to the value of the deflnition of output participation factors given in (3.16) evaluated 41 for ^A and ^C, where ^A = TAT?1 ^C = CT?1; where T is the diagonal transformation matrix required to transform the sym- metric n-ellipsoid or n-rectangle set ? into a an n-cube or an n-sphere. Figure 3.1 depicts such a diagonal transformation T for the case of a two-dimensional system. 2. The sum of output participation factors associated with any mode is equal to 1, i.e., nX i=1 CkriliCkT kCkk2 = 1 3. The output participation factors are dimensionless quantities that are inde- pendent of the units in which state and/or output variables are measured. Next, a numerical example is presented to show that the values of the output participation factors computed according to the expression in (3.10) do not agree with the values of output participation factors (deflned in (3.16)). 42 T z1 z2 x2 x1 T z1 z2 x2 x1 (a) (b) Figure 3.1: Illustrating diagonal transformation T in two dimensions, (a) Elliptic original uncertainty set ?, (b) Rectangular original uncertainty set ?. Numerical example Consider an LTI system given by (3.6){(3.7) with A = 0 BB BB BB @ 0 1 0 ?1:5 ?0:5 1:225 ?0:433 0 0 1 CC CC CC A and C = 0 BB @ 0 1 1 0 2 1 1 CC A The initial conditions of this system are uncertain and lie in a set ?. We assume that ? is an n-sphere of radius c0, i.e., kx0k ? c0 (c0 is a positive constant). Thus, no transformation is required and the output participation factors can be computed using A and C according to the deflnition in (3.16). The eigenvalues of A are 43 ?1;2 = ?0:0672?j1:2026, ?3 = ?0:3655. We are interested in the participation of the critical mode (corresponding to the complex conjugate pair of eigenvalues ?1;2) in the states and in the outputs. Right and left eigenvectors corresponding to ?1 that satisfy the normalization condition (2.2) are, respectively, r1 = ? ?0:0348?j0:6216 0:7499 0:2224?j0:0249 ?T l1 = ? ?0:0078+j0:8297 0:6087+j0:2255 0:1944?j0:6308 ? : The participations of mode 1 in states xi; i = 1;2;3 are p11 = 0:5160 ? j0:0240 (jp11j = 0:5166), p21 = 0:4565+j0:1691 (jp21j = 0:4868) and p31 = 0:0275?j0:1451 (jp31j = 0:1477). The participation of mode 1 in the outputs y1 and y2 can be calcu- lated using (3.16): py11 = 0:3854?j0:2071 (jpy11j = 0:4375) and py21 = 0:4854?j0:0690 (jpy21j = 0:4903). A simple calculation shows that participation of mode 1 in either output is not the weighted sum of the state participations (as suggested by the ex- pression given in (3.10)). Next, we present another notion of modal participation factors. This new deflnition is one of input-to-output participation factors. It in- volves probe input signals as well as modes and outputs, and is applicable to a large set of systems. 44 3.3 Input-to-Output Modal Participation Factors Consider a linear time-invariant system with inputs and outputs: _x(t) = Ax(t)+Bu(t) (3.18) y(t) = Cx(t): (3.19) The sought deflnition of input-to-output participation factors is considered a natural extension to the deflnition of input-to-state participation factors [40] given in Section 2.1.2. Recall that the deflnition of input-to-state participation factors assumes a special form of the input and output matrices B and C, respectively. It is assumed that the input is applied to one component, say the q-th component, of the right side of (3.18) and only one state, say the k-th state, is measured. For cases of systems where availability of measurements of states is not usually possible, and/or inputs are applied to more than one component of the right side of (3.18), extending the deflnition in (2.10) to address such cases is needed. We therefore, consider the case of systems with general Bq and Ck, where Bq and Ck are the q-th column of input matrix B and k-th row of output matrix C, respectively. This means that the input is applied through the q-th input channel of the right side of (3.18) and the k-th output of (3.19) is measured. With this choice of the input location and the measured output, in steady state the k-th output in (3.19) (in the frequency domain) is given by 45 yk(s) = Ck(sI ?A)?1Bqu = nX i=1 CkriliBq s??i u(s) = nX i=1 Rikq s??iu(s) Let Rikq := CkriliBq: Take pyiqk = jRikqj (3.20) as the participation factor of mode i in output k when the input is applied through the q-th input channel. We call this quantity the input-to-output participation factor (IOPF) for mode i, with measurement at output k and input applied through the q-th input channel. Note that the IOPF, similar to the ISPF, is dimensionless. Also, note that, the quantity Rikq = CkriliBq is numerically equal to the entry on the k-th row and q-th column of the residue matrix associated with mode i. This deflnition of input-to-output participation factors is useful in designing monitoring systems to detect impending instability in nonlinear systems as discussed in the next chapter. 46 Chapter 4 Detecting Impending Instability in Nonlinear Systems In this chapter, a signal-based approach for real-time detection of impending insta- bility in nonlinear systems is considered. The main idea pursued here involves using a small additive white Gaussian noise (AWGN) as a probe signal and monitoring the spectral density of one or more measured states or outputs for certain signatures of impending instability. Input-to-output participation factors introduced in Chapter 3 along with Input-to-state participation factors given in [40] are used as a tool to aid in selection of locations for probe inputs and outputs or states to be monitored. Since these participation factors are model-based, the chapter combines signal-based and model- basedideastowardachievingarobustmethodologyforinstabilitymonitoring. Power systems will be the main application of this methodology where the detection of impending instabilities is crucial. 47 The chapter proceeds as follows. In Section 4.1, the sought signal-based approach to instability monitoring using input-to-state and/or input-to-output par- ticipation factors is presented. A stability index based on power spectral density measurements is also given. In Section 4.2, three case studies from power systems are given to demonstrate the proposed approach to instability monitoring. 4.1 Precursor-Based Monitoring In this section, we show that noisy precursors can be used as a warning signal indicating that a system is operating dangerously close to instability. We also show that the spectrum of a measured output (state) of the system is proportional to the square of the input-to-output (input-to-state) participation factors. Thus, IOPFs (ISPFs) can be used to determine the best location for applying the probe signal and for choosing which output (state) to measure where the noisy precursor would be most apparent. A stability index based on power spectral density measurements that can be used to aid operators in assessing proximity to the approaching instability is given. Figure 4.1 shows a schematic diagram of our instability monitoring system. The Figure depicts the idea of having the monitoring system to trigger a preventive control action to protect the system against the efiects of the approaching instability. One possible preventive action is discussed in Chapter 5. 48 Measurements Control Power System Probe Signal Inputs Instability Detector Signal Analysis/ Preventive Figure 4.1: Precursor-based instability monitor with external probe signal. 4.1.1 Developed monitoring system Consider a nonlinear dynamic system (\the plant") _x = f(x;?)+?(t) (4.1) where x 2 Rn, ? is a bifurcation parameter, and ?(t) 2 Rn is a zero-mean vector white Gaussian noise process [25]. Let the system possess an equilibrium point x0. For small perturbations and noise, the dynamical behavior of the system can be described by the linearized system in the vicinity of the equilibrium point x0. The linearized system corresponding to (4.1) with a small noise forcing ?(t) is given by _x = Df(x0;?)x+?(t) (4.2) 49 where Df(x0;?) represents the gradient of f(x;?) with respect to x evaluated at x = x0. In (4.2), x now denotes x ? x0 (the state vector referred to x0). For the results of the linearized analysis to have any bearing on the original nonlinear model, we must assume that the noise is of small amplitude. The noise ?(t) can occur naturally or can be injected using available controls. We consider the case where the noise is applied through one input channel and the power spectral density of one output is calculated. That is, we consider the case where ?(t) = Bq?(t) where Bq is the q-th column of the input matrix B, ?(t) is a scalar white Gaussian noise with zero mean and power 2, applied through the q-th input channel and the output is given by y = Cpx where Cp is the p-th row of the output matrix C. In steady state, the p-th output of system (4.2) forced by a small AWGN is given by yp(s) = nX i=1 CpriliBq s??i ?(s) = nX i=1 Ripq s??i?(s) = H(s) ?(s): The absolute value of Ripq was deflned in Chapter 3 as the participation factor of mode i in output p when the input is applied through the q-th input channel. The power spectral density (PSD) of the output of a linear system with transfer function 50 H(j!) is related to the PSD of the input by [24] Sy(!) = H(j!)H(?j!)S?(!) Thus, the power spectrum of the p-th output is given by Syp(!) = ? nX i=1 Ripq j! ??i !? nX k=1 Rkpq ?j! ??k ! 2 = 2 nX i=1 nX k=1 Ripq j! ??i Rkpq ?j! ??k Suppose that the system is nearing a Hopf bifurcation. Speciflcally, assume that a complex conjugate pair of eigenvalues is close to the imaginary axis, and has rela- tively small negative real part in absolute value compared to other system eigenval- ues. Denote this pair as ?1;2 = ???j!c, with ? > 0 small and !c > 0: jRe(?i)j ?; i = 3;:::;n: Under this assumption, and for values of ! close to !c, Syp(!) can be approximated as Syp(!) ? 2 2X i=1 2X k=1 Ripq j! ??i Rkpq ?j! ??k = 2 jR 1pqj2 ?2 +(! ?!c)2 + jR1pqj2 ?2 +(! +!c)2 + 2Re ? 1 ?+j(! ?!c) (R1pq)2 ??j(! +!c) ? : Note that all terms containing ?i, i = 3;:::;n have been neglected and only terms containing the critical eigenvalues ?1 and ?2 have been retained. After algebraic manipulation and substituting (R1pq)2 = fi + jfl where fi = jR1pqj2 cos(2 pq) and 51 fl = jR1pqj2 sin(2 pq), with pq = tan?1 ? ImfR1pqg RefR1pqg ? , the PSD of yp can be rewritten as Syp(!) = 2jR1pqj2 1 ?2 +(! ?!c)2 + 1 ?2 +(! +!c)2 ? + 2(fl?+fi!c)(! ?!c)+?(?fi?!cfl)(?2 +!2 c)(?2 +(! ?!c)2) ? 2(fl?+fi!c)(! +!c)??(?fi?!cfl)(?2 +!2 c)(?2 +(! +!c)2) = 2jR1pqj2 ? (1+G1(!)) 1?2 +(! ?! c)2 + (1?G2(!)) 1?2 +(! +! c)2 ? (4.3) where G1(!) = 1?2 +!2 c [(?sin(2 pq)+!c cos(2 pq))(! ?!c) + ?(?cos(2 pq)?!c sin(2 pq))]; G2(!) = 1?2 +!2 c [(?sin(2 pq)+!c cos(2 pq))(! +!c) ? ?(?cos(2 pq)?!c sin(2 pq))]: For ! = !c and su?ciently small ? (? ? !c), the power spectral density of yp is given by Syp(!c) = 2jR1pqj2 1 ?2 +O 1 ? ? +O(1) ? : Note that the IOPFs are related to the spectral densities of the outputs of a system driven by small AWGN as in (4.3). The amplitude of the spectrum is proportional to the square of the IOPFs. The input-to-output participation factors can be used 52 to determine the best location for applying the probe signal and also the output that will have the highest spectral peak. The special case where only system states are monitored and input is applied to one component of the right side of (4.2) bears similar results [21]. In [21], we show that the ISPFs are related to the spectral densities of the states of a system driven by small AWGN. The amplitude of the spectrum is proportional to the square of the ISPFs. Thus, the input-to-state participation factors can also be used to determine the best location for applying the probe signal and also the state that will have the highest spectral peak. 4.1.2 Instability proximity index In this section, an instability proximity index that helps predict closeness to insta- bility based on power spectral density measurements is given. Performance indices have been used to predict proximity to voltage collapse in power systems [44]. Sev- eral stability indices have been proposed in the literature, some of these indices are model-based while others rely mainly on online measurements and operators expe- rience. Sensitivity factors, singular values and eigenvalues are among those indices used in the literature [44]. In this paper, a sensitivity-based index is proposed. This index is based on online measurements of power spectral density peaks at certain critical frequencies. 53 The proposed index is given by PSDI = dfPSDvgd? : (4.4) Here, PSDI stands for power spectral density index, PSDv is the power spectral density peak value at the critical frequency of the measured variable v (state or output) corresponding to some value of the bifurcation parameter ?. The PSDI value at time t is approximated using ?fPSDvg=?f?g at time t, where ?fPSDvg and ?f?g are the changes in the values of PSDv and ? between time instances t and t ? ?t, respectively. The sampling time ?t is determined based on oper- ator?s experience and system?s conditions. The instability index is calculated for the variable v (state or output) which has the highest ISPF or IOPF, respectively. The instability index grows signiflcantly as the system approaches criticality, where it is theoretically inflnity at criticality. The reciprocal of the index is more infor- mative than the index itself as the reciprocal value approaches zero as the system approaches criticality. This index can be used to assist a system operator or an automatic controller in taking a preventive action. For example, a threshold can be used such that if the reciprocal of the index drops below that threshold, a preventive action is triggered. The efiectiveness of the proposed instability index in detecting closeness to instability in a single generator power system model is demonstrated in Section 4.2.2. 54 4.2 Power System Examples In this section, we apply results from the previous section to monitor and detect impending instability of examples from power systems. These examples show how ISPF and IOPF can be used to better monitor an impending instability. Although white Gaussian noise was used to derive the results presented so far, band-limited white Gaussian noise signals will be used to probe the monitored power system examples. This is because it is practically impossible to generate white noise signals. Band-limited white noise is a useful theoretical approximation when the noise disturbance has a correlation time that is very small relative to the natural bandwidth of the system [43]. 4.2.1 Single generator with dynamic load Consider the single generator power system model with induction motor load [12]: _?m = ! (4.5) M _! = ?dm! +Pm ?EmVYm sin(?m ??) (4.6) Kqw _? = ?Kqv2V 2 ?KqvV +Q(?m;?;V)?Q0 ?Q1 (4.7) TKq!Kpv _V = Kp!Kqv2V 2 +(Kp!Kpv ?Kq!Kpv)V + Kq!(P(?m;?;V)?P0 ?P1) ? Kp!(Q(?m;?;V)?Q0 ?Q1) (4.8) 55 The state variables are ?m (the generator phase angle, closely related to the me- chanical angle of the generator rotor), ! (the rotor speed), ? (the load voltage phase angle) and V (the magnitude of the load voltage). The load includes a constant PQ load in parallel with an induction motor. The real and reactive powers supplied to the load by the network are P(?m;?;V) = ?E00VY 00 sin(?)+EmVYm sin(?m ??); Q(?m;?;V) = E00VY 00 cos(?)+EmVYm cos(?m ??) ? (Y 00 +Ym)V 2 where E00 = E0p1+C2Y ?2 0 ?2CY ?1 0 cos 0 Y 00 = Y0 q 1+C2Y ?20 ?2CY ?10 cos 0 00 = 0 +tan?1 ? CY ?1 0 sin 0 1?CY ?10 cos 0 The values of the parameters for this model are: M = 0:01464, C = 3:5, Em = 1:05, Y0 = 3:33, 0 = 0, m = 0, Kp! = 0:4, Kpv = 0:3, Kq! = ?0:03, Kqv = ?2:8, Kqv2 = 2:1, T = 8:5, P0 = 0:6, P1 = 0:0, Q0 = 1:3, E0 = 1:0, Ym = 5:0, Pm = 1:0, dm = 0:05. All values are in per unit except for angles, which are in degrees. It has been shown that a supercritical Hopf bifurcation occurs in this power system model as the reactive load Q1 is increased through the critical value Q?1 = 2:9801438 [12]. Next, we consider the system operating at loads close to the Hopf 56 bifurcation, say at Q1 = 2:9. The corresponding operating point is x0 = [0:2473, 0; 0:0398; 0:9248]. The Jacobian of the system at this operating point is A = 2 66 66 66 66 66 4 0 1 0 0 ?324:5254 ?3:4153 324:5254 ?73:8611 33:3333 0 ?29:2479 72:7220 ?3:3656 0 1:5180 ?11:1529 3 77 77 77 77 77 5 The eigenvalues of A are f?0:7923?j6:6318; ?21:1157?j10:9959g. To monitor the system, a band-limited AWGN probe signal is applied to the mechanical power Pm. Figure 4.2 depicts the spectral densities for the four states ?m, !, ? and V for Q1 = 2:9 and = 0:001. As it is clear from this flgure, the state ! has a higher peak than all other states. Figure 4.3 demonstrates the variation of the spectral density peak near ! = !c ? 6:6 rad=s as a function of the bifurcation parameter Q1. The values of the input-to-state participation factors of the critical mode in all states are given in Table 4.1. As predicted by the analysis in Section 4.1, the ordering of the peaks of the spectral densities of all states at !c can be predicted from the values of the ISPFs. 4.2.2 Single generator connected to an inflnite bus Consider a synchronous machine connected to an inflnite bus together with excita- tion control [14]. It was shown in [14] that this system undergoes a Hopf bifurcation as the control gain in the excitation system is increased beyond a critical value. The 57 0 5 10 15 20 25 30 35 40 10?7 10?6 10?5 10?4 10?3 10?2 10?1 Freq [rad/s] Power spectral density ?m ?? V Figure 4.2: Power spectral densities of the states of the model given in (4.5)-(4.8). The bifurcation parameter was set to Q1 = 2:9. Band-limited AWGN of zero mean and (0:001)2 power was added to Pm. 58 2.5 2.6 2.7 2.8 2.9 30 0.02 0.04 0.06 0.08 0.1 0.12 Q1 Value of PSD of ? at critical freq ? c Figure 4.3: Variation of the peak value of the power spectral density of ! as a function of the bifurcation parameter Q1. Band-limited AWGN of zero mean and (0:001)2 power was added to Pm. dynamics of the generator are given by: _? = ! (4.9) 2H _! = ?D! +!0(Pm ?Pe) (4.10) ?0d0 _E0q = EFD ?E0q ?(Xd ?X0d)id (4.11) 59 States Spectral peak Input-to-state participation at !c ? 6:6318 factors (ISPFs) ?m 9:528?10?4 p121 = 2:8992 ! 40:38?10?4 p122 = 19:364 ? 6:616?10?4 p123 = 2:4013 V 0:305?10?4 p124 = 0:4981 Table 4.1: Input-to-state participation factors and spectral peaks at !c. with the following algebraic equations: Pe = Eqiq Eq = E0q +(Xq ?X0d)id id = x(Eq ?Ecos?)?rEsin? iq = r(Eq ?Ecos?)+xEsin? x = Xl+XqR2 l+(Xl+Xq)2 r = RlR2 l+(Xl+Xq)2 The subscripts d and q refer to the direct and quadrature axes, respectively. The dynamics of the excitation control are given by ?E _EFD = ?KEEFD +VR ?EFDSE(EFD) (4.12) ?F _V3 = ?V3 + KF? E (?KEEFD +VR ?EFDSE(EFD)) (4.13) ?A _VR = ?VR +KA(VREF ?Vt ?V3) (4.14) Here Vt is the terminal voltage and is given by V 2t = v2d +v2q 60 where ?vd = ?q = ?Xqiq vq = ?d = E0q ?X0did: The saturation function SE(EFD) is usually approximated as SE(EFD) = AEX exp(BEXEFD): An equilibrium point of this system is denoted by x0 = (?0;!0;E0q0;E0FD;V 03 ;V 0R). The values of the parameters that appear in this power system model are given in Table 4.2. Synchronous machine Exciter Transmission line H = 2:37 s KE = ?0:05 R0l = 0:02 D = 1 pu KF = 0:02 X0l = 0:40 Xd = 1:7 ?E = 0:50 s Rl = ?R0l X0d = 0:245 ?F = 0:60 s Xl = ?X0l Xq = 1:64 ?A = 0:10 s !0 = 377:0 rad/s AEX = 0:09 ?0d0 = 5:9 s BEX = 0:50 Table 4.2: Parameter values for the single generator connected to an inflnite bus model. For Pm = 0:937, VREF = 1:130, and ? = 2, it has been shown that a subcritical Hopf bifurcation occurs at K?A = 193:74 [14]. Next, we consider the system operating before the Hopf bifurcation, say at KA = 185. The corresponding operating point is given by x0 = [1:3515, 0, 1:1039, 2:3150, 0, 0:5472]T. The Jacobian of the system at this operating point is 61 A = 2 66 66 66 66 66 66 4 0 1 0 0 0 0 ?62:2 ?0:2 ?79:7 0 0 0 ?0:2 0 ?0:4 0:2 0 0 0 0 0 ?1:1 0 2 0 0 0 0 ?1:7 0:1 125:9 0 ?1157:6 0 ?1850 ?10 3 77 77 77 77 77 77 5 : The eigenvalues of A are f?0:0139 ? j7:7707; ? 4:5832 ? j12:6178;?2:1029 ? j0:9417g. Note that for this model, there are two physically feasible locations for ap- plying the probe signal. The probe signal can be either applied to Vref or to Pm. The input-to-state participation factors are used to determine the best location for applying the probe signal. From the values of the ISPFs (see Table 4.3), it is clear that mode 1 has higher participation in other states when the probe signal is applied to Pm than when applied to Vref. This can be also seen from the power spectral densities shown in Figures 4.4 and 4.5. Also, the ISPFs give an indication of which state to monitor. The higher the participation factor of the critical mode in a state, the higher the peak of the spectrum for that state. Figure 4.6 shows the variation of the power spectral peak at the critical frequency as a function of the bifurcation parameter when noise is added to Pm. Figure 4.7 shows the inverse of the PSDI based on measurements from the state VR versus the exciter gain KA. Note that, for example, a threshold value of 5 can be used such that if the reciprocal of the index drops below that value, a preventive action is triggered. 62 State Spect. peak at !c ISPFs Spect. peak at !c ISPFs (noise added to Pm) (noise added to Vref) ? 0.0226 p121 = 0:0648 0.0019 p161 = 0:0024 ! 1.2880 p122 = 0:4923 0.1084 p162 = 0:0185 E0q 0:75938?10?4 p123 = 0:0038 0:68651?10?5 p163 = 0:0001 EFD 0.2326 p124 = 0:2084 0.0210 p164 = 0:0078 V3 2:4644?10?4 p125 = 0:0068 2:2288?10?5 p165 = 0:0003 VR 3.3923 p126 = 0:8006 0.3064 p166 = 0:0301 Table 4.3: Input-to-state participation factors and spectral peaks at !c ? 7:8 for the single generator connected to an inflnite bus system. 4.2.3 Three-generator nine-bus power system Below, we consider the Western System Coordinating Council (WSCC) 3{machine, 9{bus power system model, which is widely used in the literature [33, pp. 170{ 177],[16]. The dynamics of this model includes three identical IEEE-Type I exciters for the three machines. The machines and exciters data are given in Tables 4.4 and 4.5, respectively [33, 16]. In this model, a subcritical Hopf bifurcation occurs as the load on bus 5 is increased beyond 4.5 pu [33]. Our goal in this case study is to detect this impending loss of stability by using a band-limited AWGN probe signal and continuously mon- itoring the power spectral densities of certain states or outputs. This would give the system operator (or an automatic controller) valuable time to take appropriate preventive measures (e.g., shedding loads at certain buses or trigger a controller 63 2 4 6 8 10 12 14 10?10 10?8 10?6 10?4 10?2 100 Freq [rad/sec] Power spectral density ?? Eq?E FDV 3V R Figure 4.4: Power spectral densities of the states of the single generator connected to an inflnite bus system. The bifurcation parameter was set to KA = 185. Band- limited AWGN of zero mean and (0:000032)2 power was added to Pm. that assigns the participation factors of system states in the critical mode as will be discussed in Section 6.2). The simulations of this model were conducted us- ing the software package PSAT [16]. For values of the load on bus 5 close to 4.0 pu, the linearization of the system at the operating point has two complex con- jugate pair of eigenvalues close to the imaginary axis, ?1;2 = ?0:17665 ? j8:184 and ?3;4 = ?0:3134?j1:7197. As the load on bus 5 is increased further, the pair ?3;4 approaches the imaginary axis, while the other pair ?1;2 changes only slightly. For example, when the load at bus 5 is 4.4 pu, ?1;2 = ?0:18231 ? j8:0978 and ?3;4 = ?0:04602?j2:1151. Increasing the load on bus 5 beyond 4.5 pu causes the pair ?3;4 to cross the imaginary axis from left to right. 64 2 4 6 8 10 12 14 10?8 10?6 10?4 10?2 100 Freq [rad/sec] Power spectral density ?? E?qE FDV 3V R Figure 4.5: Power spectral densities of the states of the single generator connected to an inflnite bus system. The bifurcation parameter was set to KA = 185. Band- limited AWGN of zero mean and (0:000032)2 power was added to Vref. From the values of the ISPFs calculated for this system, we found that both of the critical modes have higher participation when the probe signal is applied to Pm3, the mechanical power of generator number 3. Also, we found that these modes have high participation in the fleld voltage of the exciters. Therefore, in the following simulations, the probe signal is added to Pm3 and the power spectral densities of the fleld voltages of the three exciters (i.e., Efdi;i = 1;2;3) are monitored. Table 4.6 shows input-to-state participation factors for the 3-machine nine-bus system (partial listing) when the load at bus 5 is 4.4 pu. Figure 4.9 and Figure 4.10 show the power spectral densities of Efdi, i = 1;2;3 when the load on bus 5 (PL5) is 4.0 pu and 4.4 pu, respectively. It is clear from Figure 4.9 that when the load on bus 5 is 65 170 175 180 185 1900 0.2 0.4 0.6 0.8 1 1.2 1.4 KA Value of PSD of V R at critical freq ? c Figure 4.6: Variation of the peak value of the power spectral density of VR as a function of the bifurcation parameter KA. Band-limited AWGN of zero mean and (0:000032)2 power was added to Pm. 4.0 pu, the spectrum has two peaks at 0.28 Hz and 1.3 Hz. These two frequencies correspond to the complex eigenvalues ?3;4 and ?1;2, respectively. Note that the peak at 1.3 Hz that corresponds to the pair of complex eigenvalues ?1;2 is higher than the peak at 0.28 Hz. However, when the load at bus 5 is increased to 4.4 pu, the peak at 0.28 Hz becomes much larger than the one at 1.3 Hz (see Figure 4.10), which is an indicator that an instability is being approached. Figure 4.11 shows the power spectral density of Efd1 for three values of PL5: 4.0 pu, 4.25 pu and 4.4 pu. Next, we consider using IOPFs to decide which is the best output to mea- sure. Table 4.7 shows partial listings of the IOPFs when the load at bus 5 is 4.4 66 178 180 182 184 186 188 190 192 1940 10 20 30 40 50 60 70 80 Bifurcation parameter Ka 1/PSDI Thershold value=5 A Figure 4.7: Plot of the inverse of the power spectral density index of ! versus the bifurcation parameter KA. The threshold value is equal to 5. pu. The values listed correspond to cases where the probe signal is applied to one of the mechanical powers of the generators fPm1; Pm2;Pm3g and one of the outputs fV5;V6;V7;Q14;Q39;Q27g is measured. fV5;V6;V7g are voltages at buses f5;6;7g, respectively and fQ14;Q39;Q27g are reactive powers owing between buses f(1;4);(3;9);(2;7)g, respectively. Figure 4.12 shows the power spectral density for state Efd1 and outputs Q14 and Q27. The load value is 4.4 pu. The noise is added to Pm3, the mechanical power of generator number 3. It is clear from the flgure that IOPFs can be used to aid in the selection of signals to monitor the system other than signals of state variables. 67 Parameters Machine 1 Machine 2 Machine 3 H(secs) 23.64 6.4 3.01 xd(pu) 0.146 0.8958 1.3125 x0d(pu) 0.0608 0.1198 0.1813 xq(pu) 0.0969 0.8645 1.2578 x0q(pu) 0.0969 0.1969 0.25 T0do(sec) 8.96 6.0 5.89 T0qo(sec) 0.31 0.535 0.6 Table 4.4: Three-generator nine-bus power system: machine data. Parameters KA TA(sec) KE TE(sec) KF TF(sec) 20 0.2 1.0 0.314 0.063 0.35 Table 4.5: Three-generator nine-bus power system: exciter data. States measured Input noise Efd1 Efd2 Efd3 !1 !2 !3 added to Pm1 3.0017 2.6973 2.1357 0.0033 0.0028 0.0031 Pm2 2.6113 2.3465 1.858 0.0029 0.0024 0.0027 Pm3 4.7816 4.2967 3.4022 0.0052 0.0044 0.0049 Vref1 0.0155 0.014 0.0111 0.0000169 0.0000143 0.000016 Vref2 0.0233 0.021 0.0166 0.0000255 0.0000215 0.00002409 Vref3 0.0475 0.0427 0.0338 0.0000519 0.000043 0.000049 Table 4.6: Input-to-state participation factors for the 3-machine nine-bus system (partial listing). The load at bus 5 is 4.4 pu. 68 230 KV Gen 1 230 KV Gen 3 13.8 KV 1.025 pu Gen 2 18 KV 1.025 pu 90 MW 30 MVAR 125 MW 50 MVAR Gen 1 16.5 KV 1.04 pu 1 4 5 6 8 9 7 2 3 85 MW Slack Bus 100 MW 35 MVAR 230 KV 163 MW Figure 4.8: WSCC 3-machine, 9-bus system. Outputs measured Input noise V5 V6 V7 Q14 Q39 Q27 added to Pm1 0.8255 0.6490 0.6922 4.2022 0.9882 1.4460 Pm2 0.7182 0.5646 0.6021 3.6557 0.8597 1.2579 Pm3 1.3152 1.0339 1.1027 6.6948 1.5743 2.3037 Vref1 0.0043 0.0034 0.0036 0.0218 0.0051 0.0075 Vref2 0.0064 0.0050 0.0054 0.0327 0.0077 0.0112 Vref3 0.0131 0.0103 0.0110 0.0666 0.0157 0.0229 Table 4.7: Input-to-output participation factors for the 3-machine nine-bus system (partial listing). The load at bus 5 is 4.4 pu. 69 0 0.5 1 1.5 2 10?6 10?5 10?4 10?3 10?2 10?1 Freq [Hz] Power spectral density Efd1Efd 2Efd 3 Figure 4.9: Power spectral densities of the states Efd1, Efd2 and Efd3. The load on bus 5 was used as a bifurcation parameter. The load value is 4.0 pu. Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. 70 0 0.5 1 1.5 2 10?5 10?4 10?3 10?2 10?1 Freq [Hz] Power spectral density Efd1Efd 2Efd 3 Figure 4.10: Power spectral densities of the states Efd1, Efd2 and Efd3. The load on bus 5 was used as a bifurcation parameter. The load value is 4.4 pu. Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. 71 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 10?5 10?4 10?3 10?2 10?1 Freq [Hz] Power spectral density of E fd 1 PL 5 =4.0 pu PL 5 =4.25 pu PL 5 =4.4 pu Figure 4.11: Power spectral density of Efd1 for three values of PL5 (the load on bus 5): PL5 = 4:0 pu (dash-dotted line), PL5 = 4:25 pu (dashed line) and PL5 = 4:4 pu (solid line). Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. 72 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.610?6 10?5 10?4 10?3 10?2 Freq [Hz] Power spectral density Q27Q 14Efd 1 Figure 4.12: Power spectral densities of the state Efd1 and outputs Q14 and Q27. The load on bus 5 was used as a bifurcation parameter. The load value is 4.4 pu. Band-limited AWGN of zero mean and 0:05 power was added to Pm3, the mechanical power of generator number 3. 73 Chapter 5 Participation Factor Assignment Using Eigenvector Assignment-Based Feedback Control In this chapter, we develop state feedback control to assign modal participation fac- tors for systems nearing instability using eigenvector assignment-based techniques. We focus on assigning participation factors for state variables within a predeflned group of state variables, where the group can be any subvector of the state vector, or could consist of all system states. A procedure for computing the desired closed- loop right eigenvector(s) (based on a given desired closed-loop participation factors) is presented. This chapter proceeds as follows. In Section 5.2, the algorithms for assign- ing participation factors of system states in a mode associated with a single real eigenvalue approaching zero are presented. In Section 5.3, the case of assigning 74 participation factors for systems with an open-loop mode associated with a pair of conjugate complex eigenvalues approaching the imaginary axis is discussed. 5.1 Introduction The goal of the control efiort is not to shift the eigenvalue(s) associated with a mode that is approaching the imaginary axis (denoted as the critical mode). Rather, the goal is to assign the participation factors of a predeflned group of states in the critical mode. We would typically desire that some of these participation factors be small, and possibly that others be relatively large so that the energy in the mode afiects these physical states more than the \protected" ones. This could be viewed as using some states, which could be associated with devices inserted into the system, as vibration absorbers that tend to shield the most valued parts of the system from the instability especially in its initial stages. The feedback control will be designed to only change the closed-loop right eigenvector(s) associated with the critical mode. The remaining closed-loop right eigenvectors (associated with the remaining eigenvalues of the system) will be kept the same as their open-loop values. We address two cases, flrst the case of assigning participation factors of states in a mode associated with a single real eigenvalue approaching zero, and second the case of assigning participation factors of states in a mode associated with a pair of conjugate complex eigenvalues approaching the imaginary axis. 75 Consider the closed-loop LTI system _x(t) = (A+BF)x(t) (5.1) associated with an open-loop system _x(t) = Ax(t)+Bu(t) (5.2) to which a state feedback u(t) = Fx(t) is applied. In Chapter 4, it has been shown that noisy precursors can be used as a warning signal indicating that a system is operating dangerously close to instability. This warning signal can be used to trigger some control system as the one presented in this chapter. The objective will be to move the system to a safer operating condition in the sense that the \high-value" vulnerable parts of the system are more \protected". Minimizing the participation factors of a high-value group of states in the critical mode implies that some other group of states will experience higher magni- tudes of participation factors in the critical mode. Deciding which group of states will be seen as \low-value" is based on system conditions and operators experience. To avoid confusion, we will denote the group of high-value and low-value states as the control group states. Also, we will arrange the system states such that high-value states are followed by low-value states followed by the rest of the system states. Based on the number of states in the control group and the number of system inputs m, assigning participation factor associated with a critical mode 76 can be divided into two cases, (i) when nng ? m, and (ii) when nng > m, where nng is the number of states in the control group. In the flrst case, we will show that exact assignment of participation factors in the control group is possible. In the second case, an optimal solution will be sought as the freedom of exact assignment is limited by the number of system inputs (which is less than the number of states in the control group). 5.2 Participation Factor Assignment: Case of Real Eigenvalue Approaching Zero In this section, we develop feedback control algorithms for assigning participation factors of states in a mode associated with a single real eigenvalue approaching zero. Such a mode is considered critical in the sense that it is associated with the possibility of the system losing stability. Indeed, in such cases, the system would typically undergo some type of stationary bifurcation at the critical parameter values where the zero eigenvalue is actually attained. Next, we show how to compute the desired closed-loop right eigenvector ri0d based on a given (desired) speciflcation of the closed-loop participation factors. Note that the superscript (0) is used to denote closed-loop quantities. 77 5.2.1 Computing desired closed-loop right eigenvector ri0d The desired closed-loop i-th right eigenvector ri0d can be written as ri0d := 2 66 66 66 4 rs rns rnc 3 77 77 77 5 ; where, rs 2 Rns; rns 2 Rnns; rnc 2 Rn?(ns+nns) represent vectors of components of ri0d corresponding to high-value, low-value and uncontrolled states, respectively. fns; nns; nncgrepresentthenumberofhigh-value, low-valueanduncontrolledstates, respectively. The high-value and low-value states form a control group. Let nng = ns +nns be the number of control group states. The case of zero closed-loop participation factors of high-value states in mode i is straightforward and can be achieved by setting rs = 0 and rns = e 2Rnns where e is a column vector of nonzero elements (to avoid zero participation). The remaining (n?nng) components are left unspecifled (unconstrained). Next, we consider the general case of nonzero closed-loop participation fac- tors. Let the desired closed-loop participation factors of the control group states fx2;:::;xnngg in mode i be given in terms of the closed-loop participation factor of state x1 (flrst state of the group), i.e., let zk, k = 2;:::;nng be the desired ratio between the closed-loop participation factor of state xk and that of state x1. Next, we state a proposition that focuses on solving this participation factor assignment problem in the case where we only desire to attain a certain ratio of closed-loop 78 participation factors for two states. The proposition shows how to compute the ratio between components of the closed-loop right eigenvector ri0 (associated with the real i-th eigenvalue ?i) to achieve a given (desired) ratio between the closed-loop participation factors of two states fxk; xjg in mode i. Proposition 5.1 Consider a linear time-invariant system (5.1{5.2) with the goal of achieving a desired ratio z between the closed-loop participation factors of any two states fxk; xjg in the i-th mode (viewed as the critical mode). It is possible to achieve this desired ratio through state feedback without changing any right eigenvector except for ri. Moreover, this is achieved by assigning the k-th and j-th components frik0; rij0g of the closed-loop i-th right eigenvector such that their ratio is zli k=l i j , where lik; lij are the k-th and j-th components of the open-loop i-th left eigenvector, respectively. Proof: See Appendix A.1. For the general case where more than two participation factors are to be assigned, let zk, k = 2;:::;nng be the desired ratio between the closed-loop par- ticipation factor of state xk and that of state x1. The flrst nng components of the closed-loop i-th right eigenvector should be selected such that the ratio between the k-th component and the flrst component is zkli k=l i1, for k = 2;:::;nng. The remaining (n?nng) components are unspecifled (unconstrained). 79 5.2.2 Feedback control design: exactly assignable case In this case, the number of states in the control group (formed of high-value and low-value states) is less than or equal to the number of inputs m, i.e., nng = ns +nns ? m: Recall the main assumption of the assignment problem, that is, the controller should not change any open-loop eigenvalue or any open-loop right eigenvector except the eigenvector associated with mode i. Saying that, observe that for all open-loop pairs (?j; rj), fj = 1;2;:::;n; j 6= ig, conditions (1) and (2) of Proposition 2.1 are trivially satisfled. Moreover, by deflnition, rj lies in the null space of (?jI?A), and hence in the span of N?j (satisfying condition (3) of the Proposition). Therefore, according to Preposition 2.1, there exists a feedback gain matrix F such that all open-loop pairs (?j; rj), fj = 1;2;:::;n; j 6= ig can be kept invariant from their original, uncontrolled values. Next we recall the steps to compute the state feedback gain matrix F given in Section 2.3.1 with appropriate modiflcations to solve the assignment problem beforehand. Steps to compute F 1. Using the procedure given in Section 5.2.1, compute the desired vectors rs; rns of components of the i-th closed-loop right eigenvector ri0 (to be exactly as- signed) corresponding to high-level and low-level states, respectively. 80 2. Form a column vector e 2Rm?nng of nonzero elements. 3. Compute k to satisfy 2 66 66 66 4 rs rns e 3 77 77 77 5 = ~N?i k (5.3) where ~N?i is the m?m matrix of the flrst m rows of N?i (deflned in (2.19)) associated with the flrst m states.1 The closed-loop right eigenvector ri0 is then given by ri0 = N?i k 4. Let wi = ?M?i k; where M?i is given by (2.19). Construct the real matrices W 2 Rm?n and V 2Rn?n W = [0 ::: 0 wi 0 ::: 0] V = [:::Refrjg Imfrjg :::ri0:::rq :::] where rj and rq represent open-loop right eigenvectors associated with complex and real eigenvalues, respectively. 1In case ~N? i is singular, a state from the control group has to be replaced by a state from outside the group and test again for the singularity of the new ~N?i. 81 5. The feedback gain matrix F is given by F = WV ?1: Numerical example Consider an example of a linear time-invariant system (5.2) ([23], pp. 108), where A = 2 66 66 66 66 66 4 0 1 0 0 0 ?1:89 0:39 ?5:555 0 ?0:034 ?2:98 2:43 0:035 ?0:0011 ?0:99 ?0:21 3 77 77 77 77 77 5 and B = 2 66 66 66 66 66 4 1 0 0:76 ?1:6 ?0:95 ?0:032 0:03 0 3 77 77 77 77 77 5 : (5.4) The system?s four eigenvalues are f?0:1051; ?1:4811? 0:6239i; ?2:0127g. Since the eigenvalue f?0:1051g is the closest to the origin, let us focus on its associated participation factors for assignment through feedback. The open-loop participation factors of system states fx1; x2; x3; x4g in the flrst mode are f1.1492, -0.0653, 0.0403, -0.1241g, respectively. The ratio between the participation factors offx1;x2g is ?17:6, i.e., the participation of x1 in the flrst mode is much higher than the participation of state x2. Suppose it is desired to reduce the ratio between the participation factors of fx1;x2g to 0:5 (i.e., it is desired to push the participation of the second state in the flrst mode to be twice that of the participation of the flrst 82 state). The right modal matrix R of the open-loop system is R = 2 66 66 66 66 66 4 0:9934 ?0:4738+0:1996i ?0:4738?0:1996i 0:4447 ?0:1044 0:8262 0:8262 ?0:8951 0:0314 ?0:1738+0:0881i ?0:1738?0:0881i ?0:0221 0:0357 ?0:0730+0:0990i ?0:0730?0:0990i ?0:0213 3 77 77 77 77 77 5 : The open-loop left eigenvector components l11; l12 are f1.1568, 0.6258g, respectively. Thus the design ratio for r110=r120 should be 0:5=(1:1568=:6258) = 0:2705. Applying the steps above to compute the feedback gain matrix F yields F = 2 66 4 0:5280 0:2857 0:5846 ?1:5851 0:1578 0:0854 0:1748 ?0:4738 3 77 5 The closed-loop system matrix Ac = A+BF is given by Ac = 2 66 66 66 66 66 4 ?0:5280 0:7143 ?0:5846 1:5851 ?0:1488 ?1:9705 0:2253 ?5:10850 0:5067 0:2401 ?2:4190 0:90900 0:0192 ?0:0097 ?1:0075 ?0:1624 3 77 77 77 77 77 5 : (5.5) The eigenvalues of Ac are the same as the open-loop eigenvalues (as expected), and the closed-loop right modal matrix R0 is R0 = 2 66 66 66 66 66 4 0:2457 ?0:4738+0:1996i ?0:4738?0:1996i 0:4447 0:9083 0:8262 0:8262 ?0:8951 0:0152 ?0:1738+0:0881i ?0:1738?0:0881i ?0:0221 ?0:3382 ?0:0730+0:0990i ?0:0730?0:0990i ?0:0213 3 77 77 77 77 77 5 : 83 Theclosed-loopparticipationfactorsaref0.1389, 0.2778, 0.0095, 0.5738g. Notethat the participation of x1 in the flrst mode is half that of x2 (as desired). Also note that, although the desired goal was met, state x2 is not the state having the highest participation in the flrst mode. State x4 is the one having the highest closed-loop participation although it was not the one with the highest open-loop participation. Thus, the exact assignment methodology can precisely assign participation factors among the states in the control group but has no authority over the participation factors for states outside the control group. 5.2.3 Feedback control design: not exactly assignable case In this case, nng = ns +nns > m: Let the desired closed-loop participation of the control group states fx2;:::;xnngg in mode i be given in terms of the closed-loop participation factor of state x1 (flrst state of the group), i.e., let zk, k = 2;:::;nng be the desired ratio between the closed-loop participation factor of state xk and that of state x1. Generally, the freedom in specifying components of the closed-loop eigenvectors is limited by the number of system inputs m as noted by Liu and Paton [23]. Thus, it is not possible to exactly assign the desired ri0d (computed in Section 5.2.1) to achieve the specifled ratios zk, k = 2;:::;nng. Instead, a best possible choice for an achievable eigenvector is made. Recall the expression to compute the best achievable vector ri0a and the 84 corresponding coe?cient vector k (given in (2.26){(2.27)), k = (NT?i N?i)?1 NT?i ri0d (5.6) ri0a = N?i (NT?i N?i)?1 NT?i ri0d: (5.7) Construction of the feedback gain matrix F is straightforward following steps (4) and (5) in Section 5.2.2. 5.2.4 Feedback control design: \sacriflcial" case In this case, no prior speciflcations on the closed-loop participation factors are given except that it is required that the participation factors of the low-value states be maximized with respect to the rest of the system states. Thus, the low-value states will be \sacriflced" to \protect" the rest of the system states (including the high- value states). A trivial assumption is n?nns ? m; otherwise, it would be possible to assign zeros to all participation factors of system states except the low-value ones to achieve inflnite relative participation of low- value states with respect to the rest of the system states. The proposed setting is to maximize the ratio between the norms of two vectors. The entries of the two vectors are those of the closed-loop participation factors of low-value states and the rest of the system states, respectively. Next, we show the setup of the optimization problem. 85 For simplicity of notation, the letter i used as a subscript with the critical eigenvalue, or used as a superscript with the right/left eigenvector associated with the critical mode, is omitted whenever it is clear from the context. The subscripts ns and t are used to denote quantities that are a function of low-value and rest-of- the-system states respectively. Also, the states are reordered such that low-value states come flrst followed by the rest of the system states. The closed-loop i-th right eigenvector can be written as: r0 = 2 66 4 r0ns r0t 3 77 5 = 2 66 4 N?ns N?t 3 77 5k (5.8) where, N?ns and N?t are the matrices formed of the rows of N? associated with low-value and the rest of the system states, respectively. Deflne column vectors Pns and Pt of closed-loop participation factors of low-value states and the rest of the system states, respectively, Pns := [p01 ::: p0ns]T Pt := [p0ns+1 ::: p0n]T where, the closed-loop participation factor of state q in the real eigenvalue i, p0qi = liq0riq0, can be written in terms of open-loop quantities and riq0 as (see (A.2) in Ap- pendix A) p0qi = jRjjR0jliqriq0: 86 Deflne diagonal matrices Lns and Lt as Lns = 2 66 66 66 66 66 4 li1 li2 ... lins 3 77 77 77 77 77 5 and Lt = 2 66 66 66 66 66 4 lins+1 lins+2 ... lin 3 77 77 77 77 77 5 ; respectively. Using (5.8), Pt and Pns can be written as Pns = jRjjR0j Lns N?ns k; (5.9) Pt = jRjjR0j Lt N?t k (5.10) respectively. The solution to the optimization problem is to flnd the coe?cient vector k that maximizes the ratio between the norms of the two vectors fPns; Ptg. The statement of the problem is max Q := Pns TPns PtTPt (5.11) subject to kTk = 1 where (?)T indicates taking the transpose of a matrix or a vector. Using (5.9){(5.10), the objective function Q in (5.11) can be written as Q = k T N?nsTLT nsLnsN?ns k kT N?tTLTt LtN?t k Denote N?nsTLTnsLnsN?ns and N?tTLTt LtN?t as G and H respectively. The objective function Q can be rewritten in terms of matrices G and H as Q = k T G k kT H k 87 It is clear that G and H are symmetric matrices. If H is not positive deflnite, we may select k to drive the denominator of this objective function to zero. This would be the best case for our design; the participation factors of the low-value states in the mode i will be inflnite relative to the participation factors of the rest of the system states. More likely, however, is the case in which H is of full rank. In this case, H will have a well-deflned matrix square root (H1=2 = LtN?t), and the maximum value of the objective function Q in (5.11) is equal to the maximum eigenvalue of E := H?1=2GH?1=2: (5.12) The associated principal direction of Q, which we may denote as vmax, determines k via k = H?1=2vmax Therefore, it is possible to analytically flnd a coe?cient vector k that solves the optimization problem. Knowing the optimal coe?cient vector kopt, the closed-loop right eigenvector ri0 can be computed using (5.8). Construction of the feedback gain matrix F is straightforward following steps (4) and (5) in Section 5.2.2. Numerical example In this example, we revisit the example given in Section 5.2.2. Recall that the system has an eigenvalue equal to f?0:1051g and it is the closest eigenvalue to the origin. We focus on the case where states fx1; x2g are the high-value states while 88 states fx3; x4g are the low-value ones. No prior speciflcations on the closed-loop participation factors are given except that it is required that the participation factors of the low-value states be maximized with respect to the rest of the system states (high-value states in this example). Since the number of high-value states is two and the system has two inputs, one might think that a \good" solution is to exactly assign zeros to the closed-loop participation factors of states fx1; x2g. This is not possible as the rank of N?t (matrix formed of the rows of N? associated with the high-value states) is two, i.e., no coe?cient vector k can be found to assign zeros to the closed-loop participation factors of states fx1; x2g in the flrst mode. Thus, our objective is to assign the closed-loop participation factors of the system states to maximize the ratio of the participation factors of states fx3; x4g to states fx1; x2g. Recall that open-loop participation factors of system states fx1; x2; x3; x4g in mode one (?i = ?0:1051) aref1.1492, -0.0653, 0.0403, -0.1241g, respectively. The ratio between the norms of the vectors of open-loop participation factors of states fx3; x4g and fx1; x2g is 0:0129. Computing matrix E in (5.12) yields E = 2 66 4 1:0437 1:8279 1:8279 3:2066 3 77 5 The eigenvalues of E are f0:0013; 4:249g, and therefore, the maximum attainable value of the objective function Q is 4:249. The corresponding coe?cient vector kopt that achieves that value is [0:5107 0:8598]T. Computing the feedback gain matrix 89 F yields F = 2 66 4 0:5998 0:3245 0:6641 ?1:8007 0:1793 0:0970 0:1985 ?0:5383 3 77 5 The closed-loop participation factors of system states fx1; x2; x3; x4g in mode one are f0.0015, 0.3244, 0.0053, 0.6687g, respectively. The ratio between the norms of the vectors of closed-loop participation factors of states fx3; x4g and fx1; x2g is 4:249. This result shows a f4:249=0:0129g ? 329{fold increase between open-loop and closed-loop cases. 5.3 Participation Factor Assignment: Case of Eigenvalue Pair Approaching Imaginary Axis In this section, we develop a state feedback control for assigning participation factors for systems nearing another form of instability. The approaching instability studied here is Hopf bifurcation, where a pair of complex conjugate eigenvalues cross the imaginary axis. Let f?i, ?i+1 = ??ig be the pair of complex eigenvalues approaching the imaginary axis. As in the case of real eigenvalue approaching imaginary axis given in Section 5.2, the goal pursued by the control efiort is to achieve a prescribed set of ratios between the norms of closed-loop participation factors of states in a group denoted as the control group. This will be done without afiecting any eigenvalues or any right eigenvectors except those right eigenvectors associated with 90 mode i. Next, we show how to compute the desired closed-loop right eigenvectors fri0d; ri+10dg based on a given (desired) speciflcation of the closed-loop participation factors. 5.3.1 Computing desired closed-loop right eigenvectors fri0d; ri+10d = ri0d?g The same notation used in Section 5.2.1 to specify difierent components of the desired closed-loop right eigenvector ri0d will be adopted here, i.e., ri0d := 2 66 66 66 4 rs rns rnc 3 77 77 77 5 ; where, rs 2 Cns; rns 2 Cnns; rnc 2 Cn?(ns+nns) represent vectors of components of ri0d corresponding to high-value, low-value and uncontrolled states, respectively. fns; nns; nncgrepresentthenumberofhigh-value, low-valueanduncontrolledstates, respectively. The high-value and low-value states form a control group. Let nng = ns +nns be the number of control group states. The case of zero closed-loop participation factors of high-value states in mode iisstraightforwardanditcanbeachievedbysettingrs = 0andrns = e 2Cnns where e is a column vector of elements with nonzero norms (to avoid zero participation). The remaining (n?nng) components are unspecifled (unconstrained). For the general case of nonzero closed-loop participation factors, let the 91 desired norms of the closed-loop participation factors of the control group states fx2;:::;xnngg in mode i be given in terms of the norm of the closed-loop partici- pation factor of state x1 (flrst state of the group), i.e., let zk, k = 2;:::;nng be the desired ratio between the norm of the closed-loop participation factor of state xk and that of state x1. Next, we show how to write the closed-loop participation factor of the j-th state xj in mode i in terms of the closed-loop right eigenvector ri0. Recall that p0ji = lij0rij0 (5.13) Denote by R0 the matrix of closed-loop right eigenvectors R0 = 2 66 66 66 66 66 66 66 66 4 r11 ::: ri10 ri10? ::: rn1 ... ... ... ... ... ... r1j ::: rij0 rij0? ::: rnj ... ... ... ... ... ... r1n ::: rin0|{z} ri0 rin0?|{z} ri+10 ::: rnn 3 77 77 77 77 77 77 77 77 5 where ri0 and ri+10 = ri0? are the only closed-loop eigenvectors that are difierent from their corresponding open-loop values. The j-th component of the i-th closed- loop left eigenvector lij0, which lies in the i-th row and j-th column of the matrix of closed-loop left eigenvectors L0 (L0 = R0?1), is given by: lij0 = (?1) j+i detfR0g detfMijg; (5.14) 92 where Mij (Mij 2 C(n?1)?(n?1)) is the matrix that results from removing the j-th row and the i-th column of R0, i.e., Mij = 2 66 66 66 66 66 66 66 66 66 66 4 r11 ::: ri?11 ri10? ri+21 ::: rn1 ... ... ... ... ... ... ... r1j?1 ::: ri?1j?1 rij?10? ri+2j?1 ::: rnj?1 r1j+1 ::: ri?1j+1 rij+10? ri+2j+1 ::: rnj+1 ... ... ... ... ... ... ... r1n ::: ri?1n rin0?|{z} ri+10 ri+2n ::: rnn 3 77 77 77 77 77 77 77 77 77 77 5 Let Qj := [(?1)i+1jCi1j:::(?1)i+j?1jCi(j?1)j 0|{z} j-th (?1)i+jjCijj:::(?1)i+n?1jCi(n?1)j]T where matrices Cim (Cim 2 C(n?2)?(n?2)) are formed by removing the m-th row and i-th column of Mij, (m = 1;:::; n ? 1). The determinant of Mij can now be obtained by cofactor expansion along the i-th column, yielding an expression in terms of ri+10 = ri0? and Qj: detfMijg = ri0HQj (5.15) where (?)H indicates taking the Hermitian, or complex conjugate transpose, of a matrix or vector. Using (5.13){(5.15), the closed-loop participation factor p0ji in (5.13) can be written as p0ji = f(?1) j+i detfR0gg?fr i0HQjg?ri j 0 (5.16) 93 Recall from (2.22) that ri0 can be written as ri0 = N? k for some vector k 2 Cm and N? deflned as given in (2.18). Using that, p0ji can be expressed as p0ji = f(?1) j+i detfR0gg?fk HN?H Qjg?N? j k (5.17) where N?j is the j-th row of N?. To simplify the expression in (5.17), denote by ?j the m?m matrix ?j := N?H Qj N?j: (5.18) Equation (5.17) now becomes p0ji = (?1) j+i detfR0g ?k H?j k (5.19) Note that square matrix ?j (?j 2C(m?m)) is singular and of rank one. Next, we state a proposition that focuses on solving this participation fac- tor assignment problem in the case where we only desire to attain a certain ratio of closed-loop participation factors for two states. The proposition shows how to compute the i-th closed-loop right eigenvector ri0 (associated with the complex i-th eigenvalue ?i) to achieve a given (desired) ratio between the norms of closed-loop participation factors of two states fxj; xlg in mode i. Proposition 5.2 Consider a linear time-invariant system (5.1{5.2), with the goal of achieving a desired ratio z between the norms of the closed-loop participation 94 factors of any two states fxj; xlg in the critical mode \i." Suppose that the matrices f?j; ?lg don?t have the same null space. (Recall that matrices f?j; ?lg are given by (5.18).) Then, it is possible to achieve the desired ratio z of participation factors without afiecting any right eigenvectors except those associated with mode i. Proof: See Appendix A.2. Thus, accordingtoProposition5.2, itislikelytoflndvectorskj, j = 2;:::;nng that solve for all desired ratios zj, j = 2;:::;nng. Despite that, it might be impos- sible to flnd a single vector k that simultaneously solves for all desired ratios zj, j = 2;:::;nng. A candidate coe?cient vector kcad could be a weighted sum of the computed kj?s that solves for the zj?s. Those weights can be chosen based on the most desired ratios. If for example, the ratios are equally desired, then the candidate vector kcad is simply the average of the computed kj?s. 5.3.2 Feedback control design: exactly assignable case Unlike exactly assignable cases discussed in Section 5.2.2 for systems with a real eigenvalue approaching zero, the cases discussed in this section are limited. Two cases are discussed, case (a) where we only desire zero closed-loop participation of high-value states (of dimension less than number of inputs m) in mode i, and case (b) where we only desire to attain certain ratio between norms of closed-loop participation factors for two states only. The feedback control design follows a similar approach to the one given for the real eigenvalue case except for some minor 95 changes. Recall the main assumption of the assignment problem, that is, the controller should not change any open-loop eigenvalue or any open-loop right eigenvector ex- cept the eigenvectors (ri; ri+1 = ri?) associated with mode i. Saying that, the same argument presented in the real eigenvalue case holds, i.e., according to Proposition 2.1, there exist a feedback gain matrix F such that all open-loop pairs (?j; rj), fj = 1;2;:::;n; j 6= fi;i + 1gg can be kept invariant from their original, uncon- trolled values. Next we recall the steps to compute the state feedback gain matrix F given in Section 2.3.1 with the appropriate modiflcations to solve the assignment problem beforehand. Steps to compute F: case (a) 1. Set rs (vector of closed-loop right eigenvector components corresponding to high-value states) to zeros. 2. Form a column vector e 2Cm?ns of elements of nonzero norms. 3. Compute k to satisfy 2 66 4 rs e 3 77 5 = ~N?i k where ~N?i is the m?m matrix of the flrst m rows of N?i (deflned in (2.18)) associated with the flrst m states.2 The closed-loop right eigenvector ri0 is 2In case ~N? i is singular, see footnote in page 81. 96 then given by ri0 = N?i k 4. Let wi = ?M?i k: Construct the real matrices W 2Rm?n and V 2Rn?n W = [0 ::: 0 Refwig Imfwig 0 ::: 0] V = [:::Refrjg Imfrjg :::Refri0g Imfri0g:::rq :::] where rj and rq represent open-loop right eigenvectors associated with complex and real eigenvalues, respectively. 5. The feedback gain matrix F is given by F = WV ?1: Steps to compute F: case (b) 1. Using the result stated in Proposition 5.2, compute the coe?cient vector k that achieves the desired ratio of participation factors between two states. The closed-loop right eigenvector ri0 is then given by ri0 = N?i k 2. Follow steps (4) and (5) of the procedure to compute F given for case (a) above. 97 5.3.3 Feedback control design: not exactly assignable case This case involves the assignment of closed-loop participation factors of more than two states. Assuming the existence of solution vectors kj, j = 2;:::;nng that solve for all desired ratios zj, j = 2;:::;nng, a candidate coe?cient vector kcad is a weighted sum of the computed kj?s that solves for the zj?s (as discussed in Section 5.3.1). The weights are chosen based on system?s operation and operator?s experience. The closed-loop right eigenvector ri0 is given by ri0 = N?i kcad: Construction of the feedback gain matrix F is straightforward following steps (4) and (5) of the procedure to compute F given for case (a) of the exactly assignable case. Next, we present a numerical example of participation factor assignment involving more than two states. Numerical example The example in Section 5.2.2 is revisited but with a slight modiflcation. The system considered here has an extra input, i.e., the input matrix B has an extra column, B = 2 66 66 66 66 66 4 1 0 2 0:76 ?1:6 ?5 ?0:95 ?0:032 3 0:03 0 2:8 3 77 77 77 77 77 5 : 98 As noted in the example in Section 5.2.2, The system?s four eigenvalues are f?0:1051;?1:4811 ? 0:6239i;?2:0127g. For the mode associated with the pair of complex conjugate eigenvalues (eigenvalues 2 and 3), let us focus on its associ- ated participation factors for assignment through feedback. The open-loop par- ticipation factors of system states fx1; x2; x3; x4g in that mode are f?0:0446 ? 0:1059i; 0:0129 + 0:1872i; 0:3868 ? 1:1663i; 0:6449 + 1:0850ig, respectively. The norms of the open-loop participation factors are f0:1149; 0:1877; 1:2287; 1:2621g. Suppose that three states fx1; x2; x4g are selected to form the control group with state x4 chosen as the high-value state and states fx1; x2g are the low-value ones. The open-loop participation ratios are jp12jjp42j = 0:0910 and jp22jjp42j = 0:1487. Note that the open-loop participation of the high-value state x4 is higher than the participation of the low-value states fx1; x2; x4g in the second (and third) mode. Suppose it is desired to assign the participation of states fx1; x2; x4g to achieve the following desired ratios between norms of closed-loop participations: zd1 = jp012jjp0 42j = 1:2 and zd2 = jp022jjp0 42j = 2:24. Using the procedure given in the proof of Proposition 5.2 (see Appendix A), the coe?cient vectors fk1; k2g that separately solve for the desired ratios fzd1; zd2g, respectively, are k1 = [0:3299 ?0:1856+0:3824i 0:8452+0:3967i]T; k2 = [0:3531 ?0:1558+0:3824i 0:7623+0:3967i]T Choosing kcad to be the average of fk1; k2g yields, kcad = [0:3415 ?0:1707+0:3824i 0:8037+0:3967i]T 99 Table 5.1 shows the difierent possibly achievable closed-loop ratios fz1; z2g for difierent coe?cient vectors fk1; k2g and for the candidate vector kcad. z1 = jp012jjp0 42j z2 = jp022jjp0 42j k1 1.20 1.84 k2 1.45 2.24 kcad 1.31 2.02 Table 5.1: Closed-loop ratios fz1; z2g for coe?cient vectors fk1; k2; kcadg. The desired closed-loop ratios are fzd1 = 1:2; zd2 = 2:24g. 100 Chapter 6 Application to Electric Power Networks In this chapter, we apply the results obtained in Chapter 5 on participation factor assignment for LTI systems to an example from electric power networks. The system (at equilibrium) is operating close to instability where a pair of complex eigenvalues is close to the imaginary axis. Modal analysis of linearized model of that network is performed. Generators highly afiected by the critical mode are identifled. State feedback control is applied to assign the participation factors of a group of the state variables to protect high-value generating units. This chapter proceeds as follows. In Section 6.2, Western System Coordinat- ing Council (WSCC) 3{generator, 9{bus power system model, is examined. This system has a critical electrical mode associated with the system exciters. A state feedback control is designed for assigning closed-loop participation factors. The goal fulfllled is to trade the risk of participation in the critical mode between a high-value generator and the other generators in the system denoted as low-value generators. 101 In Section 6.3, remarks on the use of state feedback in power networks are given. 6.1 Introduction For small disturbance analysis in electric power networks, the participation factor of a mode in a state can be interpreted, from the view point of probability, as the participation \risk". Larger norm of participation factor of a state in a particular mode means it has more efiect on and will be more afiected by this mode. It may be desirable that the participation factor of a particular state variable be reduced in some critical system modes. This desire is based on the fact that the higher the magnitude of participation factor the higher the tendency of such state to be afiected in a serious way by an impending (or true) instability. In power systems operation, nuclear units, for example, are regarded as high value subsystems. Reducing par- ticipation factors of these units in some critical mechanical and/or electrical mode would reduce the risk of premature tripping, as well as alleviate the efiect on these units of a troublesome mode being excited. 6.2 Three-Generator Nine-Bus Power System In this section, we revisit the Western System Coordinating Council (WSCC) 3- generator, 9-bus power system model, studied for monitoring impending instabilities in Section 4.2.3 (see Figure 4.8 for the line diagram). In this model, a subcritical 102 Hopf bifurcation occurs as the load on bus 5 is increased beyond 4.5 pu [33]. It is observed that the critical modes associated with the pair of complex conjugate eigenvalues closest to the imaginary axis are the electrical ones associated with the exciter. The system is composed of three synchronous generators equipped with three identical IEEE-type I exciters. A fourth order model is used to capture the dynamics of the synchronous generator: _? = ?b(! ?1) _! = (Tm ?Te ?D(! ?1))=M _e0q = (?fs(e0q)?(xd ?x0d)id +v?f)=T0d0 _e0d = (?e0d +(xq ?x0q)iq)=T0q0; where the subscripts d and q refer to the direct and quadrature axes, respectively. fs(:) captures the saturation characteristic of the synchronous generator. The elec- trical power is Pe = (vq +raiq)iq +(vd +raid)id and current link is described by the equations: 0 = vq +raiq ?e0q +(x0d ?x1)id 0 = vd +raid ?e0d +(x0q ?x1)iq 103 The exciter is depicted in Figure 6.1 and described by the following equations: _vm = (V ?vm)=T _vr1 = (Ka(Vref ?vm ?vr2 ? KfT f vf)?vr1)=Ta (6.1) vr = 8 >>> >>>< >>>> >>: vr1 if vrmin ? vr1 ? vrmax ? vr1 vrmax ifvr1 > vrmax; vrmin ifvr1 < vrmin: _vr2 = ?(KfT f vf +vr2)=Tf _vf = ?(vf(1+Se(vf))?vr)=Te The generators and exciters data are given in Section 4.2.3. Details of loads and Figure 6.1: Exciter for the three-generator, nine-bus system. transmission lines can be found in [33]. The system has 24 states arranged as follows: x(t) = [?1 !1 e0q1 e0d1 vm1 vr11 vr21 vf1 :::]T: 104 When the load at bus 5 is 4.4 pu, the linearization of the system at the operating pointhasacomplex conjugatepair ofeigenvalues?fi;i+1g = ?0:04602?j2:1151. This isthepairofcomplexconjugateeigenvaluesclosesttotheimaginaryaxis. Denotethe mode associated with that pair of eigenvalues as the critical mode. Computing the participation factors of system states in that mode reveals that states representing q- axis transient voltage, e0q, are the states with the highest participation in the critical mode. Table 6.1, shows the states having the highest open-loop participation factors in the critical mode. Note that, the norms of the participation factors were divided by the sum of the norms of the participation factors to get more sense out of the numbers for comparison. State jriqj \riq jpqij = jliqjjriqjx50n k=1jlikjjrikj e0q3 0.0108 66.802 0.19887 e0q1 0.0189 34.437 0.14365 e0q2 0.0230 43.401 0.09643 e0d1 0.0339 9.702 0.07081 e0d2 0.0323 5.327 0.05138 vr21 0.0481 -59.185 0.05207 vr22 0.0433 -55.0678 0.07031 Table 6.1: The states with the highest norms of open-loop participation factors in the critical mode associated with ?fi;i+1g = ?0:04602?j2:1151. The load at bus 5 is 4.4 pu. Observe that e0q3, q-axis transient voltage of the third generator, is the state with the highest participation in the critical mode indicating that the third gen- 105 erator is the most afiected generator by the critical mode. Assume that the third generator is a high-value generator (nuclear unit for example). We would like to lower the participation of this generator in the critical mode to alleviate any unde- sirable efiects that may possibly occur if the critical mode is excited. The critical mode of the system can be revealed from the time domain simulation by applying a 2% perturbation to the initial condition of the q-axis transient voltage of generator 2 (e00q2) while keeping the rest of the states at their equilibrium values. The resulting q-axis transient voltage of generators 1; 2, and 3 are shown in Figure 6.2. From Figure 6.2, it is obvious that after short-time transients of the three generators, a poorly damped mode with frequency around 0:33 Hz can be detected (note the sus- tained oscillations). This is consistent with that the critical mode is associated with a pair of eigenvalues with frequency of 0:3337 Hz. Note also how the 3 states reacts difierently to the perturbation. e0q3, for example, is the most afiected followed by e0q2 followed by e0q1. This is expected from the values of the norms of the participation factors of the q-axis transient voltage states in the critical mode (see Table 6.1). The open-loop ratios between the norms of the participation factors of states fe0q1, e0q2g and state e0q3 are f0:7223; 0:4849g, respectively. Suppose it is desired that the closed-loop ratios between the norms of the participation factors of states fe0q1, e0q2g and state e0q3 are fz13 = 2:92; z23 = 2:03g, respectively. Particularly, we need to trade the risk of participation in the critical mode between the third genera- tor (high-value generator) and the other two generators (low-value generators) of the 106 system. Next, we design a state feedback control to assign the system participation factors in the critical mode to fulfll the goal mentioned above. 0 10 20 30 40 50 60 70 1.142 1.143 1.144 1.145(a) e q 3? 0 10 20 30 40 50 60 70 0.86 0.865 0.87 0.875 (b) e q 2? 0 10 20 30 40 50 60 700.92 0.922 0.924 0.926 Time (sec) e q 1? (c) Figure 6.2: Open-loop response for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) of the WSCC 3-gen., 9-bus power system. (a) e0q3, (b) e0q2 and (c) e0q1. 6.2.1 Feedback control design Recall the procedure for the control design given in Section 5.3.3. We have two desired ratios fz13; z23g relating the closed-loop norms of the participation factors of q-axis transient voltage states, e0qk; k = 1; 2; 3, to achieve. The control action will take place through the the reference input of the exciter connected to each generator. The input matrix B is a matrix of dimension (24?3) of zeros except at 107 entries (14;1); (18;2) and (22;3) corresponding to input vref of generators 1; 2 and 3, respectively (see (6.1)). The nonzero entries are identical and equal to KaTa = 100. Using the procedure given in the proof of Proposition 5.2 (see Appendix A), the coe?cient vectors fk1; k2g that separately solve for the desired ratios fz13; z23g, respectively, are k1 = [0:2753+0:1238i 0:5075?0:4070i 0:1508?0:5426i]T; k2 = [0:2347?0:0288i 0:6427?0:1942i ?0:0374?0:5528i]T: Choosing kcad to be the average of fk1; k2g yields, kcad = [0:2550+0:0475i 0:5751?0:3006i 0:0567?0:5477i]T: Table 6.2 shows the difierent possibly achievable closed-loop ratios for difierent co- e?cient vectors fk1; k2g and for the candidate vector kcad. z13 z23 k1 2.92 1.227 k2 1.9425 2.03 kcad 2.4314 1.6341 Table 6.2: Closed-loop ratios fz13; z23g for coe?cient vectors fk1; k2; kcadg. The desired closed-loop ratios are fz13 = 2:92; z23 = 2:03g. Table 6.3 shows the states having the highest closed-loop participation factors in the critical mode. As expected, the participation of the q-axis transient voltage state of generator 3 in the critical mode is reduced, and it is not the state with the 108 State jpqij = jliqjjriqjx50n k=1jlikjjrikj e0q3 0.0864 e0q1 0.20754 e0q2 0.13934 e0d1 0.07036 e0d2 0.04882 vr12 0.05019 vr21 0.07605 Table 6.3: The states with the highest norms of closed-loop participation factors in the critical mode associated with ?fi;i+1g = ?0:04602?j2:1151. The load at bus 5 is 4.4 pu. highest participation in the critical mode. Time domain simulations of this model were conducted using PSAT [16]. Figures 6.3-6.5 show the difierent time response of the q-axis transient voltage states, fe0q1; e0q2; e0q3g, for a 2% perturbation to the initial condition of the q-axis transient voltage of generator 2 (e00q2) before and after control. The Figures show the improvement in the response of the high-value state e0q3 of the third machine and the efiects of increased participation in the critical mode of the low-value states fe0q1; e0q2g of the flrst and second machine, respectively. 109 10 20 30 40 50 60 70 0.918 0.919 0.92 0.921 0.922 0.923 0.924 0.925 0.926 0.927 Time (sec) e q 1? Before controlAfter control Figure 6.3: Plot of q-axis transient voltage of generator 1 (e0q1) versus time for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) before and after control. 110 10 20 30 40 50 60 70 0.855 0.86 0.865 0.87 0.875 Time (sec) e q 2? After controlBefore control Figure 6.4: Plot of q-axis transient voltage of generator 2 (e0q2) versus time for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) before and after control. 111 10 20 30 40 50 60 701.141 1.1415 1.142 1.1425 1.143 1.1435 1.144 1.1445 1.145 1.1455 Time (sec) e q 3? Before controlAfter control Figure 6.5: Plot of q-axis transient voltage of generator 3 (e0q3) versus time for a perturbation of 2% in the initial condition of the q-axis transient voltage of generator 2 (e00q2) before and after control. 6.3 Remarks on Use of State Feedback in Power Networks The work in this chapter assumes access to the system state vector for purposes of feedback control. There are several ways to deal with this assumption from a practical perspective. One approach is to use a dynamic observer to calculate the state vector from available output measurements. Another approach is to calculate a state feedback control and then approximate it by an output feedback controller. The third approach is to actually make real-time state measurements available, and this is a real possibility in light of recent technological developments in real-time 112 phasor measurements [45]. As noted by Phadke [45], synchronized phasor measure- ments provide the flrst real possibility of providing a dynamic state estimator. Using such measurements, it is possible for the substations to provide a continuous stream of phasor data to the control center. In this way, a state vector estimate that follows the system dynamics can be constructed. With normal dedicated communication circuits, a continuous data stream of one phasor measurement every 2-5 cycles (33.3 - 83.33 msec) can be sustained. Considering that typical power system dynamic phenomena fall in the frequency range of 0-2 Hz, it is possible to observe power system dynamic phenomena in real-time with high fldelity at the control center. 113 Chapter 7 Conclusions and Future Directions 7.1 Conclusions In this dissertation, we have developed monitoring and control systems to improve the performance of systems that are sometimes required to operate at the edge of their stability envelopes. Our main contributions are as follows: ? We used the concept of modal participation factors extensively in this work, and generalized the concept in several directions. In particular, we introduced output participation factors and input-to-output participation factors. ? A signal-based approach for real-time detection of impending instability in nonlinear systems was considered. The main idea pursued involves using a small additive white Gaussian noise as a probe signal and monitoring the spec- tral density of one or more measured states or outputs for certain signatures of impending instability. Input-to-state and input-to-output participation fac- 114 tors were introduced as tools to aid in selection of locations for probe inputs and states or outputs to be monitored, respectively. Since these participation factors are model-based, the work presented combines signal-based and model- based ideas toward achieving a robust methodology for instability monitoring. We also introduced a proximity index to help assess closeness to instability based on power spectral density measurement. ? We also introduced the problem of participation factor assignment by feed- back, and made important progress on this problem. Feedback algorithms were developed for assigning modal participation factors using eigenvector assignment-based techniques. The goal is to reduce the interaction between a selected group of states (high-value group) and an undesirable mode. In partic- ular, we addressed two cases, flrst the case of a mode associated with a single real eigenvalue approaching zero (Stationary bifurcation), and second the case of a mode associated with a pair of conjugate complex eigenvalues approach- ing the imaginary axis (Hopf bifurcation). A novel procedure for computing the desired closed-loop right eigenvector(s) associated with the critical mode (based on given constraints on the desired closed-loop participation factors) was presented ? We applied our results to a well-studied electric power network, the WSCC 3-generator, 9-bus network. As expected, participation factor assignment has a clear efiect on the relative time domain responses of system states. 115 7.2 Future Directions Among the many open problems of interest are the following: ? More work is needed on generalized participation factors and on participation factor assignment by state and output feedback. Particularly, the problem of flnding a deflnition for participation of outputs in modes that reduces to the original state participation when the output is simply one of the states is yet to be solved. Unlike state participation problem that has identical expressions for participation of modes in states and states in modes, the output participation problem is quite difierent and an expression for participation of outputs in modes is not expected to be the same as the expression proposed for participation of modes in outputs in Chapter 3. ? A particularly challenging problem involves detection not only nearness to in- stability, but also detecting the severity of the impending instability from the point of view of nonlinear system behavior. For example, an oscillatory insta- bility can be of the hunting type, in which small amplitude oscillations occur, or it can be divergent, resulting in complete loss of operation. Although this can be determined using analytical models using known methods of bifurca- tion analysis, it is not known how this can be achieved using a signal-based approach. ? Another direction involves studying use of other probe signals besides AWGN. 116 Examples include periodic signals, chaotic signals covering an appropriate fre- quency range, and colored noise signals. The relative advantages and disad- vantages of the various probe signals should be considered. In this regard, connections to past work in real-time probing of power systems and aircraft dynamics should be studied. In research aircraft, for example, it is common to use \chirp" signals to probe the aircraft for its stability properties in various parts of its ight envelope. 117 Appendix A Proof of Propositions 5.1 and 5.2 A.1 Proof of Proposition 5.1 The matrix of closed-loop right eigenvectors R0 is given by R0 = [r1 ::: ri0 ::: rn]; where ri0 (associated with the real eigenvalue ?i) is the only closed-loop right eigen- vector afiected by the control law. The matrix of closed-loop left eigenvectors L0 = R0?1 is given by L0 = [l10T ::: li0T ::: ln0T]T: The closed-loop i-th left eigenvector li0 is given by li0 = 1detfR0g[ci1 ::: cik ::: cin]; where the cik?s, k = 1;2;:::;n, represent matrix cofactors, i.e., cik = (?1)k+i ?detfMikg; 118 where Mik is the matrix of dimension (n?1) that is obtained by eliminating the k-th row and the i-th column of R0. Since the only change to the open-loop right modal matrix R is to the components of its i-th column ri, the cik?s, k = 1;2;:::;n are only function in open-loop left eigenvectors lq?s, q = 1;:::; n; q 6= i. Therefore, the k-th component of the closed-loop i-th left eigenvector li0 can be written in terms of the k-th component of the open-loop i-th left eigenvector lik as lik0 = detfRgdetfR0glik: (A.1) Using (A.1), the closed-loop participation factor of state k in mode i is reduced to p0ki = lik0rik0 = detfRgdetfR0glikrik0 (A.2) Next, we use (A.2) to write the desired ratio z between the closed-loop participation factors of states fxk;xjg in mode i as z = p 0 ki p0ji = lik0rik0 lij0rij0 (A.3) = l i kr i k 0 lijrij0; (A.4) Therefore, to achieve the desired ratio z, the closed-loop right eigenvector compo- nents rik0; rij0 should be chosen such that rik 0 rij0 = z lik=lij. ? 119 A.2 Proof of Proposition 5.2 The closed-loop participation factor of the j-th state xj in mode i in terms of the closed-loop right eigenvector ri0 is given by (see (5.13)): p0ji = (?1) j+i detfR0g ?k H?j k (A.5) The ratio between the norms of the closed-loop participation factors of two states fxj; xlg in the critical mode \i" is: jp0jij jp0lij = jlij0jjrij0j jlil0jjril0j (A.6) = jk H?j kj jkH?l kj; (A.7) where matrices f?j; ?lg are given by (5.18). Matrices f?j; ?lg (2 Cm?m) are singular and of rank one. The goal is to flnd a coe?cient vector k (2 Cm) that achieves a desired ratio z between the norms of the closed-loop participation factors of the two states fxj; xlg in the critical mode \i," i,e., flnd k that solves: jp0jij jp0lij = jkH?j kj jkH?l kj (A.8) = z: In general, if there are any solutions other than 0, one would expect the solutions to form a variety of codimension 1 inCm. If we can flnd some k0 for which the right side of (A.8) is less than z and some k1 for which it is greater than z, then there is a solution k, k = t k0 +(1?t) k1; 0 < t < 1: (A.9) 120 The squares of the norms in (A.8) are quartic polynomials in t. Solving these quartic polynomials for t, we can flnd k that solves for the desired ratio. One possible choice of k0 is a vector that lies in the null space of ?j and not in the null space of ?l (in this case, the right side of (A.8) is equal to zero). One possible choice of k1 is a vector that lies in the null space of ?l and not in the null space of ?j (in this case, the right side of (A.8) is equal to inflnity). Note that this is always possible providing that the assumption on matrices f?j; ?lg of Proposition 5.2 holds. Knowing t, the coe?cient vector k that solves for the desired ratio z can be computed using (A.9). ? 121 Appendix B Table of Symbols Symbol Meaning detfAg or jAj determinant of A Refvg Real part of v Imfvg Imaginary part of v AT; xT transpose of matrix A, vector x A?; x? conjugate of matrix A, vector x AH; xH conjugate transpose of matrix A, vector x jaj absolute value (magnitude, if a is complex) of a kxk norm of vector x ej j-th unit vector (j-th entry is 1; 0 otherwise) Rn(Cn) real (complex) n-dim. space Rm?n(Cm?n) real (complex) m?n matrices 122 Bibliography [1] F. L. Pagola, I. J. Perez-Arriaga and G. C. Verghese, \On sensitivities, residues and participations: Applications to oscillatory stability analysis and control," IEEE Transactions on Power Systems, Vol. 4, No. 1, pp. 278{285, Feb. 1989. [2] I. J. Perez-Arriaga, G. C. Verghese and F. C. Schweppe, \Selective modal analy- sis with applications to electric power systems, Part I: Heuristic introduction," IEEE Transactions on Power Apparatus and Systems, Vol. 101, No. 9, pp. 3117{3125, Sep. 1982. [3] P. Kundur, Power System Stability and Control. McGraw Hill, New York, 1994. [4] C. L. DeMarco, \Design of predatory generation control in electric power sys- tems," Thirty-First Annual Hawaii International Conference on System Sci- ences HICSS, Vol. 3, pp. 32{38, Jan. 1998. [5] E. Abed, A. Tesi and H. Wang, \Control of bifurcation and chaos," in The Control Handbook, (CRC Press), 1996. 123 [6] E. H. Abed and J. H. FU, \Local feedback stabilization and bifurcation control, I. Hopf bifurcation," Systems and Control Letters, Vol. 7, pp. 11{17, February 1986. [7] E. H. Abed and J. H. FU, \Local feedback stabilization and bifurcation control, II. stationary bifurcation," Systems and Control Letters, Vol. 8, pp. 467{473, May 1987. [8] A.N. Andry, E.Y. Shapiro and J.C. Chung, \Eigenstructure Assignment for Linear Systems,"IEEE Trans. on Aerospace and Electronic Systems, Vol. 19, No. 5, pp. 711{729, Sept. 1983. [9] L. C. P. da Silva, Y. Wang, V. F. da Costa and W. Xu, \Assessment of generator impact on system power transfer capability using modal participation factors," IEE proceedings on Generation, Transmission and Distribution," Vol. 149, No. 5, pp. 564{570, Sept. 2002. [10] Y. Y. Hsu, C. L. Chen, \Identiflcation of Optimum Location of Stabilizer Ap- plications Using Participation Factors," IEE Proceedings on Generation, Trans- mission and Distribution," Vol. 134, pp. 238{44, May 1987. [11] I. J. Perez-Arriaga, L. Rouco, F. L. Pagola and J. L. Sancha, \The role of participation factors in reduced order eigenanalysis of large power systems," 124 IEEE International Symposium on Circuits and Systems, Vol. 927, pp. 923{ 927, June 1988. [12] H. O. Wang, E. H. Abed, and M. A. Hamdan, \Bifurcation, chaos, and crises in voltage collapse of a model power system," IEEE transactions on Circuits and Systems- I: Fundamental Theory and Applications, Vol. 41, No. 4, pp. 294{302, March 1994. [13] P. Huang and Y. Hsu, \Eigenstructure assignment in a longitudinal power system via excitation control," IEEE transactions on Power Systems, Vol. 5, No. 1, pp. 96{102, Feb. 1990. [14] E. H. Abed and P. P. Varaiya, \Nonlinear oscillations in power systems," In- ternational J. Electrical Power and Energy Systems, Vol. 6, No. 1, pp. 37{43, Jan. 1984. [15] E. H. Abed, D. Lindsay and W. A. Hashlamoun, \On participation factors for linear systems," Automatica, Vol. 36, No. 10, pp. 1489{1496, Oct. 2000. [16] F. Milano, \PowerSystem Analysis ToolboxDocumentationfor PSAT," version 1.3.0, 2004 (available online at http://www.power.uwaterloo.ca/~fmilano/). [17] B. C. Moore, \On the exibility ofiered by state feedback in multivariable systems beyond closed loop eigenvalue assignment," IEEE Trans. on Automatic 125 Control, pp. 689{692, Oct. 1976. [18] M. A. Hassouneh, H. Yaghoobi and E. H. Abed, \Monitoring and control of bifurcations using probe signals," in Dynamics, Bifurcations and Control, Lec- ture Notes in Control and Information Sciences, Vol. 273, F. Colonius and L. Gruene, Eds., Berlin: Springer Verlag, pp. 51{65, 2002. [19] J. F. Hauer and M. J. Beshir, \Dynamic performance validation in the western power system," Association of Power Exchanges, Kananaskis, Alberta, Oct. 2000. [20] C. G. Verghese, I. J. Perez-Arriaga and F. C. Schweppe, \Measuring state variable participation for selective modal analysis," IFAC symposium on digital control, New Delhi, India. [21] M.A. Hassouneah, M. S. Saad and E. H. Abed, \Signal-Based Instability Mon- itoring of Electric Power Systems," 16th IFAC World Congress, Prague, Czech Republic, 2005. [22] E. H. Abed, M.A. Hassouneah and M. S. Saad, \Instability monitoring and con- trol of power systems," Applied Mathematics for Restructured Electric Power Systems, J. H. Chow, F. F. Wu and J. A. Momoh, Eds., Springer, pp. 159{178, 2005. 126 [23] G. P. Liu and R. J. Patton, Eigenstructure Assignment for Control System Design. Wiley, New York, 1998. [24] C. W. Helstrom, Probability and Stochastic Processes for Engineers, Second Edition, Macmillan, 1991. [25] T. Kim and E. H. Abed, \Closed-loop monitoring systems for detecting impend- ing instability," IEEE Trans. Circuits and Systems{I: Fundamental Theory and Applications, Vol. 47, No. 10, pp. 1479{1493, Oct. 2000. [26] N. Martins and L. T. G. Lima, \Determination of suitable locations for power system stablizers and static VAR compensators for damping electromechan- ical oscillations in large scale power systems," IEEE Transactions on Power Systems, Vol. 5, No. 4, pp. 1455{1469, Nov 1990. [27] E. H. Abed, H. O. Wang and R. C. Chen, \Stabilization of period-doubling bifurcations and implications for control of chaos," Physica D, Vol. 70, pp. 154{164, Jan 1994. [28] C. A. Canizares, S. Corsi and M. Pozzi, \Modeling and implementation of TCR and VSI based FACTS controllers," in Technical Report AT-UCR 99/595, ENEL Ricerca, Milan, Italy, Dec 1999. [29] Z. Feng, V. Ajjarapu and D. J. Maratukulam, \A comprehensive approach for 127 preventive and corrective control to mitigate voltage collapse," IEEE Trans. on Power Systems, Vol. 15, No. 2, pp. 791{797, May 2000. [30] T. Kim, Noisy Precursors for Nonlinear System Instability with Application to Axial Flow Compressors. Ph.D. thesis, University of Maryland, College Park, 1997. [31] I. Genc, H. Schattler and J. Zaborszky, \Clustering the bulk power system with applications towards hopf bifurcation related oscillatory instability," Electric Power Components and Systems, Vol. 33, No. 2, pp. 181{198, Feb 2005. [32] V. M. I. Genc, \Hopf Bifurcation Related Coherent Oscillations in Electric Power Systems with a Clusterd Texture," Ph.D. Dissertation, Washington Uni- versity, Saint Louis, MO, Dec 2001. [33] P. W. Sauer and M. P. Pai, Power System Dynamics and Stability, Prentice Hall, New Jersey, 1998. [34] G. C. Verghese, I. J. Perez-Arriaga and F. C. Schweppe, \Selective modal analy- sis with applications to electric power systems, Part II: The dynamic stability problem," IEEE Trans. Power Apparatus and Systems, Vol. 101, No. 9, pp. 3126{3134, Sep. 1982. [35] N. Mithulananthan, C.A. Canizares, J. Reeve and G.J. Rogers, \Comparison 128 of PSS, SVC, and STATCOM Controllers for Damping Power System Oscil- lations," IEEE Transactions on Power Systems, Vol. 18, No. 2, pp. 786{792, May 2003. [36] H. Ghasemi, C. A. Canizares and A. Moshref, \Oscillatory stability limit pre- diction using stochastic subspace identiflcation," IEEE Transactions on Power Systems, Vol. 21, No. 2, pp. 736{745, May 2006. [37] K. Wiesenfeld, \Noisy precursors of nonlinear instabilities," Journal of Statis- tical Physics, Vol. 38, No. 5-6, pp. 1071{1097, 1985. [38] K. Wiesenfeld, \Virtual Hopf phenomenon: A new precursor of period doubling bifurcations," Physical Review A, Vol. 32, Sep. 1985. [39] C. Jefiries and K. Wiesenfeld, \Observation of noisy precursors of dynamical instabilities,"Physical Review A, Vol. 31, pp. 1077{1084, 1985. [40] H. Yaghoobi and E. H. Abed, \Optimal actuator and sensor placement for modal and stability monitoring," Proceedings of the American Control Confer- ence, pp. 3702{3707, San Diego, CA, June 1999. [41] H. Yaghoobi, Stability Monitoring and Control of Nonlinear Systems Susceptible to Oscillatory Bifurcations. Ph.D. thesis, University of Maryland, College Park, 2000. 129 [42] J. Lu, H. D. Chiang and J. S. Thorp, \Eigenstructure assignment by decentral- ized feedback control," IEEE Transactions on Automatic Control, Vol. 38, No. 4, pp. 587{594, April 1993. [43] J. F. Hauer and J. G. DeSteese,\A Tutorial on Detection and Characteriza- tion of Special Behavior in Large Electric Power Systems,", Paciflc Northwest National Laboratory Richland, Washington 99352, July 2004. [44] C. Canizares, \Voltage stability assessment: Concepts, practices and tools," Technical Report SP101PSS, Institute for Electrical and Electronic Engineers, 2002. [45] A. G. Phadke, \Synchronized phasor measurements in power systems," IEEE Computer App. in Power, Vol. 6, pp. 10{15, April 1993. 130