You are on page 1of 246

s 867 1,338 *

^| - rtoritTf 0t Ifyimtvat F t7 tciI'MriA

_EMPTY_

o tCII>MTIA

_EMPTY_

_EMPTY_

_EMPTY_

_EMPTY_

SYNCHRONOUS MACHINES Theory and Performance

General Electric Series SYNCHRONOUS MACHINES by Charles Concordia TRANSIENTS IN POWER SYSTEMS by Harold A. Peterson SERVOMECHANISMS AND REGULATING SYSTEM DESIGN, VOLUME I by Harold Chestnut and Robert W. Mayer TRAVELING WAVES ON TRANSMISSION SYSTEMS by L. V. Bewley TRANSFORMER ENGINEERING by the late L. F. Blume, A. Boyajian, G. Camilli, T. C. Lennox, S. Minneci, and V. M. Montsinger, Second Edition CIRCUIT ANALYSIS OF A-C POWER SYSTEMS, TWO VOLUMES by Edith Clarke CAPACITORS FOR INDUSTRY by W. C. Bloomquist, C. R. Craig, R. M. Partington, and R. C. Wilson PROTECTION OF TRANSMISSION SYSTEMS AGAINST LIGHTNING by W. W. Lewis MAGNETIC CONTROL OF INDUSTRIAL MOTORS by Gerhart W. Neumann POWER SYSTEM STABILITY Volume ISteady State Stability; Volume II Transient Stability; by Selden B. Crary FIELDS AND WAVES IN MODERN RADIO by Simon Ramo and John R. Whinnery MATERIALS AND PROCESSES edited by J. F. Young MODERN TURBINES by L. E. Newman, A. Keller, J. M. Lyons, and L. B. Wales; edited by L. E. Newman ELECTRIC MOTORS IN INDUSTRY by D. R. Shoults and C. J. Rife; edited by T. C. Johnson A SHORT COURSE IN TENSOR ANALYSIS FOR ELECTRICAL ENGINEERS by Gabriel Kron TENSOR ANALYSIS OF NETWORKS by Gabriel Kron MATHEMATICS OF MODERN ENGINEERING Volume I by the late Robert E. Doherty and Ernest G. Keller; Volume II by Ernest G. Keller VIBRATION PREVENTION IN ENGINEERING by Arthur L. Kimball

SYNCHRONOUS MACHINES Theory and Performance CHARLES CONCORDIA Analytical Engineering Department General Electric Company Schenectady, New York One of a series written fay Genera/ Electric authors for the advancement of engineering knowledge JOHN WILEY & SONS, INC., NEW YORK CHAPMAN & HALL, LTD., LONDON, 1951

Copyright, 1951, by General Electric Company All rights reserved. This book or any part thereof must not be reproduced in any form without the written permission of the publisher Printed in the United States of America

PREFACE The primary object of this book is to present a unified development of the fundamental circuit theory of the transient performance of synchronous machines as currently used by the engineers directly concerned with the prediction of machine performance. The material was written for a synchronous machines course that has been given at the General Electric Company for the past three years. The general equations developed are applied to the calculation of transient short-circuit currents and torques; steady-state power, torque, and current, both in synchronous operation and during starting; and the voltage disturbances occasioned by sudden application of load. Emphasis is on a more or less rigorous mathematical development and on obtaining a fundamental physical understanding of the machine so that the reader will be best equipped to extend the theory as he needs it. It is presumed that the reader is acquainted, but not necessarily familiar, with 1. The ordinary steady-state and transient theory of static circuits including the elementary law of electromagnetic induction in circuits. 2. The general physical appearance of a synchronous machine. 3. Ordinary differential equations, and operational calculus in at least one of its various forms. 4. A usual undergraduate course in round-rotor a-c rotating electric machinery, covering only the steady-state performance. 5. Symmetrical components. 6. The per-unit system of representation of machine and power system parameters. Thus, the book is intended primarily for the practicing engineer who wants to learn something about the transient theory of synchronous machines and who has heretofore been obliged to dig through the technical literature of the past twenty-five years to do so. It is not intended as a reference book, even though formulas for many specific cases can be found in it. On the contrary, it is intended to be read as a whole from the beginning. It will be evident to those familiar with the literature of synchronous machines that for the sake of unity many of the derivations and the notation, and in some cases the form of the results, have been revised.

vi PREFACE In particular, the method of deriving the general equations and the treatment of single-phase short circuits had to be considerably revised. Also, certain new material has been added. These new items are: the treatment of the double-line-to-ground short circuit, of the unidirectional components of short-circuit torque, of the starting torque, and of voltage dip on application of load. The theory presented in this book is the culmination of the work of many engineers over a period of about twenty-five years. Acknowledgment of sources can therefore be made only through the list of references and the bibliography, as they are otherwise literally too numerous to mention. However, I want to acknowledge specifically the continued encouragement and support of Mr. S. B. Crary and the contributions to the point of view made by Mr. Gabriel Kron. I must remark further my conviction that an essential factor in the achievement of such quality as this book may have is the atmosphere of a large industrial corporation, that combines the necessity for keeping in direct contact with the latest practice with the opportunity for specialization afforded by its size. CHARLES CONCORDIA March 1951

CONTENTS CHAPTER PAGE 1 PHYSICAL DESCRIPTION OF A SYNCHRONOUS MACHINE 1 2 MATHEMATICAL, DESCRIPTION OF A SYNCHRONOUS MACHINE 6 Voltage Relations, 8 Flux-Linkage Relations, 9 Inductance Relations, 10 Transformations of Equations, 13 Armature Voltage Equations, 16 The Operational Impedances, 18 Per-Unit Quantities, 20 Slip Test for xd and xq, 23 Short-Circuit Test for xd, 24 ZeroSequence Reactance, 25 Power Output, 25 Torque, 28 Summary, 30 Problems, 31 3 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION 32 The Steady-State Vector Diagram, 32 Field Flux Linkage, 34 Power Output, 36 Power-Angle Characteristics, 37 Stability, 40 Reactive Volt-Amperes, 44 Power-Angle Characteristics for Two Machines, 46 Summary, 52 Problems, 52 4 THREE-PHASE SHORT-CIRCUIT CURRENT 54 Synchronizing Currents, 58 Steady-State Components of Short-Circuit Current, 59 Short-Circuit Test, 66 Short Circuit with Armature Re sistance, 67 Field Current, 72 Summary, 74 Problems, 75 5 SINGLE-PHASE SHORT-CIRCUIT CURRENT 76 Line-to-Line Short Circuit, 76 Phase Quantities, 81 Line-to-Neutral Short Circuit, 81 Open-Phase Voltage for Line-to-Line Fault, 84 Harmonic Components of Voltage and Current, 85 Decrement Factors, 90 Field Current, 95 Summary, 97 Problems, 98 6 DoUBLE-LINE-TO-GROUND SHORT CIRCUIT AND SEQUENTIAL FAULTS . . 100 Symmetrical Components, 103 Rotor Decrement Factors, 108 Stator or Armature Decrement Factors, 109 Field Current, 112 Open-Phase Voltage, 113 Sequential Application of Faults, 114 Summary, 117 Problems, 118 7 SHORT-CIRCUIT TORQUES 119 Three-Phase Short Circuit with All Resistances Neglected, 119 ThreePhase Short CircuitEffect of Armature Resistance, 121 Three-Phase Short CircuitEffect of Rotor Resistance, 123 Discussion of ThreePhase Short-Circuit Torques, 127 Line-to-Line Short Circuit, 134 Torque, 135 Other Types of Short Circuit, 142 Harmonic Components of Line-to-Neutral Torque, 145 Unidirectional Components of Torque, 147 Unidirectional Component of Torque Due to D-C Component of Current, 157 Approximate Torque Equations, 160 Summary, 163 Problems, 164 vii

viii CONTENTS CHAPTER PAGE 8 STARTING TORQUE 165 Equivalent Circuit, 171 Relation to Approximate Torque Equation, 171 Comparison of "Exact" and Approximate Methods, 175 Average Torque (d- and g-Axis Method), 177 Field Excitation, 180 Summary, 183 Problems, 184 9 VOLTAGE DIP 185 Effect of Voltage Regulator, 191 Minimum Voltage, 193 Required Exciter Ceiling, 195 Saturation, 196 Exciter Response, 197 Voltage Recovery Time, 200 Effect of Initial Load, 200 Field Current, 201 Summary, 202 Problems, 203 APPENDIX A FOURIER SERIES FOR CURRENTS, AND FUNDAMENTAL-FREQUENCY COMPONENTS OF id AND lq, FOR DoUBLE-LINE-TO-GROUND FAULT 205 B TORQUE 212 REFERENCES 216 BIBLIOGRAPHY 217 INDEX ..........,,....,,............ 221

1 PHYSICAL DESCRIPTION OF A SYNCHRONOUS MACHINE A synchronous motor or generator consists essentially of two elements: the first to produce a magnetic field, the second ajset of armature coils in which voltages are produced by the relative motion of the two elements. In the usual modern machine the field structure rotates within a stator which supports and provides a magnetic-flux path for the armature windings. The exciting magnetic field is ordinarily produced by a set of coils (the field winding) on the moving element or rotor. Since most electric power is generated (and in the case of large blocks of power, consumed) as three-phase power, there are ordinarily three armature coils, disposed around the stator at 120 intervals so that, with uniform rotation of the magnetic field, voltages displaced 120 in phase will be produced in the coils. Two observations about this last statement are in order here. 1. It only applies without qualification to a two-pole (i.e., one pair of poles) machine. In a machine with, e.g., two pairs of poles, there must be correspondingly two complete sets of armature coils 180 apart, the three coils of each being set 60 apart. Figure 1 illustrates the disposition of the coils and the magnetic-flux paths for a two- and fourpole machine. More generally, the three coils of each set must be (120/n) degrees apart, and the sets (360/n) degrees apart, where n is the number of pairs of poles. It is usual and convenient to measure the distance between coils in "electrical degrees" where 360 electrical degrees corresponds to the angle included in one pole pair and 360 actual (or mechanical) degrees equals 360n electrical degrees. In terms of electrical degrees, then, the three coils of each set are always 120 apart. 2. The steady-state voltages produced (with balanced load) are always 120 apart in phase regardless of the speed of rotation of the field. That is, since (1/n) revolution (a displacement equal to the space occupied by one pole pair) will always correspond to one cycle of the gen1

PHYSICAL DESCRIPTION OF A SYNCHRONOUS MACHINE STATOR COIL SIDES Fio. 1. Arrangement of coils in 2- and 4-pole machines

PHYSICAL DESCRIPTION OF A SYNCHRONOUS MACHINE 3 erated voltage, i.e., the fundamental frequency will always be exactly n times the speed of rotation, and, since with constant rate of rotation the time required for the rotor to move any given distance is proportional to the distance moved, the time required for the field to move from any given position with respect to one coil to the corresponding position with respect to the next coil is just one third of a cycle or 120 electrical degrees. Of course, the machine is ordinarily connected to a three-phase bus to which voltage is also being supplied by other synchronous machines, so that this applied voltage will not correspond to the rotational speed unless the machine is running in synchronism with the rest of the machines on the system. Hence, the term "synchronous machine" means one that ordinarily runs in synchronism with other machines of the same general type, and the terms synchronous operation, out of synchronism, and out of step have meaning only for a machine connected to a system, and not for a single machine operating alone. Since in normal operation the magnetic flux produced by the field winding is rotating with respect to the stator windings and its supporting magnetic structure, voltages are produced in the iron as well as in the coils, and it is necessary to laminate the stator iron in order to break up the eddy-current paths and thus minimize the i2r losses and shortcircuiting effect which would otherwise result. The field structure or rotor on the other hand sustains principally only a constant flux and so does not have to be laminated throughout. When balanced three-phase armature currents of speed frequency are flowing, the mmf's produced by these currents tend to combine to give a resultant mmf that rotates at the same speed as the field. It is found that the best way to study the effects of this armature mmf is to resolve it into its space harmonics, upon which it may be discovered that the fundamental component rotates at rotor speed and so is stationary with respect to the field, while some of the space harmonics rotate at different speeds and so are moving with respect to the field. Since the armature coils are distributed along the stator surface so as to tend to minimize all harmonics other than the space fundamental, these harmonic effects may be regarded as secondary from the standpoint of performance. They contribute to the armature leakage reactance (i.e., to components of armature flux which do not link any of the rotor windings) and to rotor surface eddy-current losses which make it desirable to laminate at least the surface of the rotor iron whenever possible. In general, practically all machines except high-speed, two- or four-pole turbine-generators have laminated pole faces and are constructed with salient poles as shown in Fig. 2, whereas the rotors of two- and four-

4 PHYSICAL DESCRIPTION OF A SYNCHRONOUS MACHINE pole machines may be made as a single solid piece of steel. In Fig. 2 can also be seen a damper winding or amortisseur consisting usually of a set of copper or brass bars set in pole-face slots and connected together Fig. 2. Rotor for salient-pole machine .j^isE! A-^|g gHNMHj [.uuu,e*w ,n , '. T^&W m< W Fig. 3. Rotors for solid-rotor machines at the ends of the machine. This amortisseur has several useful functions: x to permit starting of synchronous motors as induction motors using the amortisseur as equivalent to the squirrel cage of an inductionmotor rotor, to assist in damping rotor oscillations, to reduce overvoltages under certain short-circuit conditions, and to aid in synchronizing the machine. 1 Superscripts refer to items in the list of references at the end of the book.

PHYSICAL DESCRIPTION OF A SYNCHRONOUS MACHINE 5 Figure 3 shows two rotors for two-pole solid-rotor turbine-generators. In this case the solid steel rotor itself serves the purpose of the amortisseur. From the brief description given above it is evident that the stator and rotor of a synchronous machine differ in these respects: The stator is more or less standardized and relatively simple in form for any type of synchronous machine and is, moreover, completely symmetrical with respect to any of the three phases. On the other hand, the rotor presents a considerable variety of forms, ranging from the simplest case of a single field winding on an otherwise symmetrical laminated rotor to a salient-pole rotor with an amortisseur having several windings, or to a solid rotor, which, although symmetrical except for the field, is still complex in that the solid steel rotor may be considered as equivalent to an amortisseur of infinitely many circuits.

2 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE We have pointed out in Chapter 1 that a synchronous machine consists of two major components, the stator and the rotor, that are in relative motion and that are rather different in structure. Regardless of this it is of course possible to write down the circuit-voltage equations simply in terms of the self- and mutual inductances of all the windings. In order to do this, we must first decide what the rotor windings are. We shall assume that the rotor magnetic paths and all of its electric circuits are symmetrical about both the pole and interpole axes as shown in Fig. 4 for a salient-pole machine. The field winding is of course separate from the others and has its axis in line with the pole axis. The amortisseur bars are all connected together in a more or less continuous mesh, but, if they are arranged symmetrically, current paths may be chosen which are also symmetrical about both the pole and interpole axis. Figure 4 shows the circuits used. The bars are numbered starting from the direct axis, which is in line with the pole axis. The direct axis circuits are then numbered Id, 2d, etc., to correspond with these bars. In the quadrature axis, which is taken as 90 electrical degrees ahead of the direct axis in the direction of normal rotor rotation, the circuits are numbered Iq, 2q, etc., starting outward from this axis. This symmetrical choice of the rotor circuits has the virtue of making all mutual inductances and resistances between direct- and quadrature-axis rotor circuits equal to zero. In some machines the amortisseur bars are not connected between poles, but even in these cases current may flow between poles at the ends of the machine through the rotor iron itself since the bars are not insulated. This lack of insulation means also that the circuit equations are only approximations to the actual case, in which some current may spread through the iron. This effect is small except where the interpole iron path is concerned, and except in turbinegenerators wherein the currents in the iron form the whole amortisseur 6

MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE 7 effect. Since the turbine-generator is so different in this respect from the salient-pole generator, it will be treated separately after we have seen how the salient-pole case turns out. All mutual inductances between stator and rotor circuits are periodic functions of rotor angular position. In addition, because of the rotor saliency, the mutual inductances between any two stator phases are also periodic functions of rotor angular position. We thus arrive at a POLE I "> > d Id dl It I j^^M-.-^-.-MNUMBERING OF ROTOR CIRCUITS FIQ. 4. Diagram of amortisseur circuits set of differential equations most of whose coefficients are periodic functions of rotor angle, so that even in the case of constant rotor speed (when the equations are linear if saturation is neglected) they are awkward to handle and difficult to solve. However it is found that, if certain reasonable assumptions are made, a relatively simple transformation of variable will eliminate all these troublesome functions of angle from the equations. The first assumption is that the stator windings are sinusoidally distributed along the air gap as far as all mutual effects with the rotor are concerned. This assumption of sinusoidal distribution of the stator windings may be justified from the standpoint that in practically all synchronous machines the windings are distributed so as to minimize all harmonics as much as is feasible.2 The principal justification comes from the comparison of performance calculated on that basis with actual performance obtained by test.

8 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE The second assumption is that the stator slots cause no appreciable variation of any of the rotor inductances with rotor angle. This assumption is evident for machines with a large number of slots per pole, but for machines with a very small number (especially an integral number) of slots it is not so evident. However, a small number of slots occurs principally in machines having a large number of poles, and thus any effect of the slots may be made to average out over the whole machine. Again the final justification comes from comparison of theory and test results. A third assumption which will be made in this book at least for the present is that saturation may be neglected. The effects of various assumptions regarding saturation will be shown later. The electrical performance of a synchronous machine may now be described by the following equations. Voltage Relations ARMATURE OR STATOR ea ~ Pta ria tb = ptb rib (1) ec = p\l/c ric where ea = terminal voltage of phase a fa = total flux linkage in phase a ia = current in phase a. Note that the direction of positive armature current is taken as opposite to what might have been expected in a static network in order to have positive current correspond to generator action. a, b, c, are the three phases lettered in the direction of rotor rotation as shown in Fig. 5 r = resistance of each armature winding, assumed to be the same for a, b, c p = the derivative operator d/dt, t = time FIELD + Wd (2) where here and in all the following equations the symbols e, t, i have the same meaning as above, the subscripts denoting the circuit in question.

FLUX-LINKAGE RELATIONS 9 DIRECT-AXIS AMORTISSEUR 0 = ptid + rudiid + ri2di2d -\ ---0 = pfad + r2idiid + r22di2d H ---- (3) etc. Here the subscripts I2d and 2ld denote mutual effects between circuits Id and 2d (see Fig. 4). It may be noted that the amortisseur circuits are resistar 2e-coupled as well as inductance-coupled and that there is no coupling between direct- and quadrature-axis circuits because of the rotor .symmetry about the direct and quadrature axes. QUADRATURE-AXIS AMORTISSEUR 0 = ptiq + rnqiiq + rl2qi2q -\ ---0 = pt2q + r2iqiiq + r22qi2g -\ ---- (4) etc. Flux-Linkage Relations ARMATURE XbbH ~ Xbcic + Xbfdifd + + xc2di2d -\ ----- h xclqilq + xc2qi2q H ---- (5) where the x's are inductances to be defined later and the subscripts refer, as before, to the circuits in question. FIELD ^/d = Xfadia ~ Xfbdtb ~ Xfcdlc + Xffdifd + Xf idild + Xf2di2d H ----- h Xfiqiiq + xf2qi2q -\ ---- (6)

10 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE DIRECT-AXIS AMORTISSEUR tld = xiadia Xlb<flb ~ Xlcdic + Xlfdifd + Xlldild + Xl2,fl2d + ' ' ' + Xldlqilq + XldZqiZq + ' ' ') 6tC. (7) QUADRATURE-AXIS AMORTISSEUR (8) + Xlq2cfi2d -\ ----- h + Z12^29 H Inductance Relations ARMATURE SELF-INDUCTANCES The self-inductance of any armature winding varies periodically from a maximum when the pole axis is in line with the phase axis to a minimum when the interpole axis is in line with the phase axis. Because AXIS OF PHASE a POLE, OR DIRECT AXIS INTERPOLE, OR QUADRATURE AXIS FIG. 5 of the symmetry of the rotor, the inductance must have a period of 180 electrical degrees and must be expressible by a series of cosines of even harmonics of angle. It will be shown that under the assumption of

INDUCTANCE RELATIONS 11 sinusoidal winding distribution only the first two terms of the series are significant, or Xaa = ZaaO + Xaa2 COS 26a (9) where 6a is the angle of the direct axis from the axis of phase a, measured in the direction of rotor rotation. Because of the sinusoidal winding distribution, current in phase a produces a space wave of mmf in the air gap which is only of fundamental span as far as the rotor is concerned. This may conveniently be broken up into components proportional to (cos 0a) and ( sin 0a) acting in the direct and quadrature axes, respectively (see Fig. 5). These components of mmf produce corresponding components of flux having space fundamental components of magnitude (<j>d = Pd cos 0a) and (4>q = Pq sin 0J where Pd and Pq are proportional to effective permeance coefficients in the direct and quadrature axes, respectively. Space-harmonic components of flux are also produced, but, since they do not link the stator, they do not concern us now. The linkage with phase a caused by this flux is then proportional to (see Fig. 5) <t>d cos 0a ^g sin 0a = Pd cos2 0a + Pq sin2 0a = .Pd + P" + Pd~PqCOS26a = A + BCOS26a 22 (10) There is also some flux linking phase a that does not link the rotor. This flux adds only to the constant term A of equation 10, and so the inductance remains of the form of equation 9. Similarly, Xbb = XaaO + Xaa2 COS 26b (11) Xee = XaaO + Xaa2 COS 20c where 06 = 6 - 120 (12) 0c = 0 + 120 ARMATURE MUTUAL INDUCTANCES To determine the form of the mutual inductance between, e.g., phases a and b, we may recognize first that there may be a component of mutual flux that does not link the rotor and is thus independent of angle. Then

12 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE considering current in phase a, the components of air-gap flux are, as before, proportional to <t>d = Pd cos 6a and <t>g = -Pg sin 6a (13) and the linkage with phase 6 due to these components is proportional to <t>d cos 6b <t>q sin 0(, = Pd cos 0a cos 6b + Pq sin 0a sin 6b = Pd cos 6 cos (6 - 120) + Pq sin 6 sin (0 - 120) -P q + - - cos 2(6 - 60) 42 = - JA + B cos 2(0 - 60) = -[A + cos 2(0 + 30)] The total mutual inductance is thus of the form, Xab = [XabO + Xaa2 COS 2(0 + 30)] Note that the variable part of the mutual inductance is of exactly the same magnitude as that of the variable part of the self-inductance and that the constant part has a magnitude of very nearly half that of the constant part of the self-inductance. Now that we know the answer, it may seem obvious from symmetry considerations that the mutual inductance ab should have a (negative) maximum when the pole axis is lined up 30 behind phase a or 30 ahead of phase b, and a (negative) minimum when it is midway between the two phases. We might also have reasonably taken a chance that higher harmonic terms could not appear in the mutual inductance since they dropped out of the selfinductance. Finally, we can write all the stator mutual inductances as Xab = Xba = ~[XabO + Xaa2 COS 2(0 + 30)] Xbc = Xcb = [XaW + Xaa2 COS 2(0 90)] (14) ab0 + aa2 COS 2(0 + 150)] ROTOR SELF-INDUCTANCES Since we are neglecting the effects of stator slots and of saturation, all the rotor self-inductances, //d, xnd, x22d, 119, etc., are constants.

TRANSFORMATIONS OF EQUATIONS 13 ROTOR MUTUAL INDUCTANCES All mutual inductances between any two circuits both in the direct axis and between any two circuits both in the quadrature axis are constant, and of course xfld = xifd, etc. Because of the rotor symmetry there is no mutual inductance between any direct- and any quadratureaxis circuit. Thus: Xflq ~ xf2q = Xldlq = xld2q = xlqfd = ^lld = xlq2d = 0, etc. (15) MUTUAL INDUCTANCES BETWEEN STATOR AND ROTOR CIRCUITS By considering current in each rotor winding in turn and recalling that only the space-fundamental component of the flux produced will link the sinusoidally distributed stator, we see that all stator-rotor mutual inductances vary sinusoidally with angle and that they are maximum when the two coils in question are in line. Thus: Xald = Xfad = Xafd COS 6 Xbfd = Xfbd = Xafd COS (0 120) Xcfd = Xfcd = Xafd COS (0 + 120) Xal d = Xiad = Xald COS 6 Xbl d - Xibd = Xald COS (6 120) (16) xcld = xUd = Xau cos (6 + 120), etc. Xalq = Xlaq = Xalq SU1 6 Xblq = XUq = Xalq SU1 (6 - 120) Xciq = xUq = -xalq sin (6 + 120), etc. Transformations of Equations Utilizing the mutual-inductance relations equations 16, we may rewrite the rotor flux-linkage equations 6, 7, and 8 as FIELD tfd = xafd[ia cos 6 + ib cos (0 120) + ic cos (6 + 120)] + Xffdifd + xfidiid + xndiid + (17)

14 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE DIRECT-AXIS AMORTISSEUR tid = xaid[ia cos 6 + ib cos (6 - 120) + ie cos (6 + 120)] + xndiid + xi2di2d -\ ---- , etc. (18) QUADRATURE-AXIS AMORTISSEUR \h9 = +xaig[ia sin 6 + ib sin (6 - 120) + ic sin (6 + 120)] + Xi2qi2q H ---- , etc. (19) The form of these equations suggests that they may be simplified by the substitution of new variables id, iq, IQ defined by the relations : s id = f ta cos 6 + ib cos (6 - 120) + ic cos (6 + 120)] iq = - f [*'a sin 6 + i b sin (6 ~ 12O) + ic sin (0 + 120)] (20) By reference to Fig. 5, or for that matter by inspection of equations 17, 18, and 19, we see that id and iq are proportional to the components of mmf in the direct and quadrature axes, respectively, produced by the resultant of all three armature currents ia, ib, and ic. The factor % is introduced so that, for balanced phase currents of any given (maximum) magnitude, the maximum values of id and iq as the phase of the currents is varied will be of the same magnitude. The maximum magnitude of any one of the phase currents under these balanced conditions will be given by V i2d + i2q and will also be the same. The current i0 is introduced since, if three currents ia, 4, and ic are to be eliminated, in general three substitute variables will be required. i0 is the conventional zero-phase-sequence current of symmetrical-component theory, and, if only i0 exists (i.e., if ia = ib = ie), equations 17, 18, and 19 show that no flux will link the rotor. By substituting the relations 20 in equations 17, 18, and 19, we obtain, for the rotor-circuit flux linkages: = ^Xafdid + Xffdifd + Xfidild. + ' '' = %Xaidid + xlfdifd + xndiid + xi2di2d -\ ---- (21) = fcalqiq + Xllqilg + Xl2qi2g H ---- , etC.

TRANSFORMATIONS OF EQUATIONS 15 Now, equations 9, 11, 14, and 16 are substituted for equations 5 for the armature flux linkages, to obtain ia = XooO^'o + Xabo(h + te) - X^ia COS 26 + Xaa2ib COS 2(6 + 30) + Xaa2ic cos 2(6 + 150) + (XafcCifd + Xaldild + Xa2di2d H ) COS 6 - (Xalqilq + Xa2qi2q H ) SHI 0 ib = Xaaolb + Xabo(ic + 4) + Xaa2ia cos 2(6 + 30) - Xaa&b cos 2(6 - 120) + c cos 2(6 - 90) (22) + (Xafdifd + Xaidild + xa2a42d H ) cos (6 - 120) - (Xalqilq + Xa2qi2q -\ ) sin (6 - 120) ic = Xaaoic + Xabo (la + ib) + Xaa2ia cos 2(6 + 150) + x^h cos 2(6 - 90) - xoo2ic cos 2(0 + 120) + (Xafdifd + Xaidild + Xa2dkd H ) cos (6 + 120) - (Xalqilq + xa2qHq H ) sin (6 + 120) In these flux-linkage equations the armature phase currents ia, ib, and ic may be eliminated in favor of the new variables id, iq, and i0, which does not, however, eliminate the trigonometric functions of rotor angle in this case. The form of the new equations suggests that a simplification can be effected by defining, similarly to id, iq, and i0, three new flux linkages id, iq, and i0. id = f [ia cos 6 + ib cos (6 - 120) + ic cos (6 + 120)] iq = -l\ia sin 6 + ib sin (9 - 120) + ic sin (5 + 120)] (23) io = \(ia + ib + ic) Now, if equations 22 are substituted in equations 23 and the proper trigonometric reductions are made, we obtain the relatively simple relations: id = (XaaO + XabO + ^Xaa2)id + Xafdifd + XaldHd + Xa2di2d H . iq = (Xaa0 + xabo %xaa2)iq + xaiqiiq + xa2ai2a + (24) io = (XaaO 2xabo)io

16 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE In equations 24, fa and \{/q may be regarded as corresponding to flux linkages in coils moving with the rotor and centered over the direct and quadrature axes, respectively. The equivalent direct-axis moving armature circuit has the self-inductance: Xd = ZooO + ZoW + f Zoo2 (25) and the equivalent quadrature-axis moving armature circuit has the self-inductance: Xq = XaaO + XabO ~ fZoo2 (26) There is also an equivalent zero-sequence axis coil which has the selfinductance: ZO = XaaO ~ ^abO (27) and which is completely separated magnetically from all the other coils. Armature Voltage Equations Finally, we can eliminate the phase quantities ia, ib, ie, and fa, \l/b, tc from equations 1, as they now occur nowhere else. This may most easily be done by denning new voltages ed, eq, and e0 in the same manner as the currents (equations 20) and flux linkages (equations 23). ed = f [ea cos 6 + eb cos (6 - 120) + ec cos (6 + 120)] eq = -f[ea sin 6 + eb sin (6 - 120) + ec sin (0 + 120)] (28) By substituting equations 1 in equations 28 and utilizing the relations 20, we obtain ed = f [cos 6pta + cos (6 - 120) ptb + cos (6 + 120) p^c] - rid eq = -f [sin 6 pfa + sin (6 - 120) p+b + sin (6 + 120) ptc] - riq e0 = pt0 ri0 (29) The bracketed expressions in equations 29 may be evaluated by differentiating the first two of equations 23, whence ptd = f [cos 6 pta + cos (6 - 120) ptb + cos (6 + 120) ptc] - I [fa sin 6 p6 + tb sin (6 - 120) p6 + tc sin (6 + 120) p6] or, by substituting \l/q from equations 23, Ptd = f [COS pta + COS (6 - 120) ptb + COS (6 + 120) ptc] + tqp6 .. J (30)

ARMATURE VOLTAGE EQUATIONS 17 and similarly, piq= - f [sin 0 pia + sin (0 - 120) pib + sin (0 + 120) pic] - idp6 (31) Thus, equations 29 become ed = Pid - iqpQ - rid eq = ppa + +dP0 ~ riq (32) eo = Pio - ri0 We note that these equations 32 are just like the original relations 1 but with the addition of generated- or speed-voltage terms iqp6 and idp6 in the direct- and quadrature-axis voltages. From a physical viewpoint our algebraic manipulations have corresponded to the specification of the armature quantities along axes fixed in the rotor and thus rotating with speed, p6, with respect to the stator axes. We should therefore naturally expect to find generated voltages as well as induced voltages produced by these rotating flux linkages. The complete set of machine-performance equations now consists of the circuit voltage equations 32, 2, 3, and 4, and the flux-linkage equations 24 and 21. At constant rotor speed these equations are linear differential equations with constant coefficients, and even with variable rotor speed they are considerably simpler than the original set of equations. The phase quantities ia, H, ic, ea, eb, ec, ia, ib, and ^c in any particular problem may be found from the substitute variables id, iq, and t'o, etc., by solving the relations 20, 23, and 28 to obtain ia = id cos 0 iq sin 0 + i0 ib = id cos (0 - 120) - iq sin (0 - 120) + i0 (33) ic = id cos (0 + 120) - iq sin (0 + 120) + to ia = id cos 0 \pq sin 0 + "A0 ib = id cos (0 - 120) - iq sin (0 - 120) + i0 (34) ic = id cos (0 + 120) - iq sin (0 + 120) + &, ea = ed cos 0 eq sin 0 + e0 eb = ed cos (0 - 120) - eq sin (0 - 120) + e0 (35) ec = ed cos (0 + 120) - eq sin (0 + 120) + e0 s

18 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE 'We have stated previously that the armature-voltage equations 32 are in a form especially suited to the solution of salient-pole synchronousmachine problems. However, in order to keep clear our concept of what we are doing, it should be pointed out that equations 32 in no way imply the existence of a salient-pole machine or even of any rotating machine. Equations 1 apply to any three coils of equal resistance, and the relations 33, 34, and 35 may be used to obtain equation 32, regardless of the nature of any other coils magnetically linked with these three coils. In the salient-pole machine case we have chosen 6 to be the angle of the direct axis of the rotor ahead of the axis of phase a, but this choice is not at all a necessary condition for the validity of equations 32. It is selected solely to simplify the flux-linkage relations. In case, for example, of a completely symmetric static network containing no capacitors, all choices of 6 lead to identical flux-linkage relations (in all cases simpler than the original relations in terms of threephase quantities) so that we could take 6 equal to anything from 6 = 0 to 6 = angle of any machine that we might later want to connect to our three coils. The Operational Impedances Since in many important problems one is interested primarily in the results as viewed from the machine armature terminals, as, e.g., in computing short-circuit currents, it is convenient to write the machine equations in a more compact form by eliminating the rotor currents. This may be done by (1) substituting the rotor flux-linkage relations 21 into the rotor-circuit voltage equations 2, 3, and 4, (2) solving these for the rotor currents in terms of the field voltage e/d and the armature currents id, iq, and (3) substituting the resulting relations in the armature flux-linkage relations 24. This may be a more or less difficult job of solving several simultaneous equations depending on the complexity of the amortisseur, but it is evident that, if we treat the derivative operator p = d/dt algebraically, as will be legitimate for many problems since all the flux-linkage relations and all the rotor-circuit voltage relations are linear, we shall arrive finally at a result of the form td = G(p)efd - xd(p)id tg = -xq(p)iq (36) where G(p), xd(p), xq(p) are operators expressed as functions of the derivative operator p. In the case of i/y, it is further evident that

THE OPERATIONAL IMPEDANCES 19 G(p) can be obtained by solving for fa as a function of efd with id = 0 and that Xd(p) may similarly be found by solving for fa as a function of id with efd = 0. We shall conform to the usual practice and call Xd(p), xq(p), and x0 the direct, quadrature, and zero-sequence axis operational impedances of the synchronous machine, even though it appears from their definitions that a more logical name would be "operational inductance." It has been stated previously that the direct- and quadrature-axis fluxes may be thought of as linking coils moving with the rotor and centered over the direct and quadrature, axes of the machine. This, together with the general form of equations 2, 3, 4, and 21, seems to suggest that at least in certain cases we should be able to regard the whole group of direct- (or quadrature-) axis circuits as representable by some sort of equivalent static electric circuit. For example, except for the mutual resistances, they are very similar to the equations of a manywinding transformer.4 In that event the calculation of xd(p) and xq(p) could be considerably simplified. However, an essential condition for the existence of a static equivalent circuit is the reciprocity of the mutual-inductance coefficients, and this condition is not completely satisfied in the present instance. That is, in equations 21 the mutualinductance coefficients between armature currents and rotor flux linkages are %Xafd, %Xaid, ~V2Xa2d, ", %Xalq, %Xa2q, ' ', but the mutual-inductance coefficients between armature flux linkages and rotor currents are xafd, xaid, xa2d, , xaiq, xa2q, That is, they are of only two-thirds magnitude and of opposite sign. This difficulty arises because of the transformation used for both current and flux linkage, which was chosen merely to keep the magnitudes of \p(t = V^ + t2q ) and of i(i = V i2d + i2q) unity for balanced unit flux linkages fa, fa, fa and currents ia, ib, and ic, respectively. It could easily have been avoided by other choices6 of transformation equations, but it seemed desirable to preserve the property of equal magnitudes. In any event, the difficulty is easily resolved by also changing over the rotor currents by a % factor, to obtain the flux-linkage relations: DIRECT AXIS ^d = Xdid + Xafdlfd + Xaldhd + Xa2dhd H fyd = Xafdid + Xffdlfd + Xfidhd + Xf2dhd H (37) iu = Xaidid + Xfidlfd + Xudhd + Xi2dhd H , etc.

20 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE QUADRATURE AXIS tq ~ Xqiq ~T" Xaigliq + Xa2qI2q ~\~ '' tiq = Xalqiq + Xnqliq + Xl2qI2q + (38) faq = Xa2qiq + XiZqllq + X22ql2q + ' ' ', 6tC. and the rotor-circuit voltage relations: DIRECT AXIS (ROTOR CIRCUITS) efd = Pt/d + Rfdlfd 0 = pt!d + Rudlid + Ri2dI2d + (39) 0 = Pt2d + Ri2dlid + R22dl2d + ' ' ', etc. QUADRATURE AXIS (ROTOR CIRCUITS) 0 = Ptlq + Rnqhq + #124/29 H 0 = Pt2q + Rl2qllq + R22qhq H , etC. These equations 37, 38, 39, and 40, together with the armature-circuit voltage equations 32, now constitute the complete set of machine-performance equations. The operational expressions of equations 36, G(p), xd(p), and xq(p), may be found from equations 37-40 in exactly the same way (and with exactly the same results) as from equations 2, 3, 4, 21, and 24. In addition, now an equivalent circuit may be used to represent and visualize these quantities,6"9 as will be shown later. All the inductances and resistances represented by upper-case symbols are three halves times the corresponding lower-case quantities of equations 2, 3, 4, 21, and 24 (e.g., Xafd = Y^afd), and all the rotor currents represented by upper-case symbols are two thirds times the corresponding lower-case currents (e.g., Ifd = %ifd). This nomenclature is followed in this chapter and in Chapter 3; but in following chapters we shall return to the use of lower-case letters, even though we shall at all times be concerned only with reactances as denned by equations 37 and 38 (i.e., the reciprocal system) and shall never again have occasion to use the nonreciprocal system. Thus it will not be necessary to use a nomenclature that distinguishes between the two systems. Per-Unit Quantities 7 Actually this multiplication of all rotor currents by % is not such a drastic procedure as it may at first glance seem, since the inductance (or reactance) and resistance coefficients of synchronous machines will usually be specified as per-unit values rather than as ohms or henrys

PER-UNIT QUANTITIES 21 anyway. The base values of armature current and voltage will ordinarily be determined by the machine rating, whereas the base values of the rotor currents are chosen so as to make the self-inductances of the armature, field, and outermost (in each axis) amortisseur circuits of about the same order of magnitude, as in the usual transformer equivalent circuit. It is not obvious in the present case, however, that base currents should be chosen inversely as the turns of each circuit, as is the case with transformers, because of the effect of the distribution of each winding in modifying its flux-producing effectiveness. The base field current may be taken as that value that will produce the same space-fundamental component of air-gap flux as is produced by base armature current id, that is, by per-unit balanced three-phase armature currents ia = cos 6 ib = cos (0 - 120) , = cos (0 + 120) Similarly, the base amortisseur current may be taken at that value that will produce the same space-fundamental component of air-gap flux as is produced by unit armature current id, when this amortisseur current flows in a direct-axis amortisseur winding of full (180) pitch. It is usually found most convenient' to use the same base value for all amortisseur currents in both the direct and quadrature axes. The two-thirds factor which had to be introduced into the rotor currents now makes its appearance simply by the fact that the effective turn ratio which must be used is calculated from the ratio of base field (or amortisseur) current to three halves times base armature phase current. One might be led to expect such a ratio from another point of view from the fact that unit id produces an air-gap, space-fundamental mmf exactly 1J^ times as big as, e.g., unit ^'a acting alone in the direct axis. This may be seen from equations 9, 10, and 11, as follows. When the direct axis (pole axis) is lined up with phase a, unit ia produces a flux: 4>d = Pd cos 6a = Pd <t>g = -P, sin 6a = 0 On the other hand, by equations 33 unit id corresponds to armature current ia = cos 0a, 4 = cos 06, ic = cos 0c which, regardless of rotor position, produce fluxes <t>d = Pd(cos2 6a + COs2 06 + cos2 6c) = f Pd and <t>q = P9(cos 0a sin 0a + cos 06 sin 06 + cos 0c sin 0c) = 0

22 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE Now let us designate as (x^d) the space-fundamental component of air-gap flux linkage produced by an armature current id. Then xad is a quantity slightly less than xd, and both the mutual inductance Xafd between field and armature direct axis and the mutual inductance Xand between the (perhaps imaginary) full-pitch, direct-axis amortisseur circuit and the armature direct axis are approximately equal to xad in per unit. In a large number of problems it has been customary to lump all the amortisseur circuits into one equivalent full-pitch circuit in each axis. In this case, the direct-axis circuits have generally been used with only a single mutual inductance xad among all three circuits. Now instead of considering the machine air-gap fluxes, which are after all not directly measurable, we may see what these definitions mean in terms of the terminal voltages. From equations 32, if the machine is running steadily at synchronous speed (p6 = 1.0) * and open circuit (id = iq = 0), then \l/g ^0 = 0 also, and e^ = 0, eq = \[/d. Also from the rotor-circuit equations 39 and 40, all the amortisseur currents lid, I2d, etc., /j9, I2q, etc., are zero and From equations 37, the direct-axis armature flux linkages are td = Xafdlfd (41) and, since eq = td, it is evident that for normal armature terminal voltage eq = 1.0, the required per-unit field current is Ifd = l/(Xafd), while the required per-unit field voltage is e/d Rfd/Xafd = rfd/xafd. Now, if we know the required no-load rated-voltage field voltage and current (neglecting saturation) in volts and amperes, we have direct relations to calculate the base field quantities. That is, if when the actual field current is //0 amperes the per-unit field current is l/Xafd, the base field current is //6 = (Xafdlf0) amperes. Similarly, if the actual field voltage is e/0 volts when the per-unit field voltage is Rfd/Xafd, the base field voltage is e/6 = (Xafdef0/Rfd) volts. On the other hand, if the base quantities are known, the per-unit machine impedances may be calculated as a A /0 and * The unit of time is that required for the rotor to move one electrical radian at synchronous speed. For example, for a normal system frequency of 60 cycles per second, the unit of time is 1/2ir60 %^^ second.

SUP TEST FOR xrf AND x, 23' Methods for calculating all the per-unit quantities from design data are discussed in detail in references 7. Slip Test for x and x If balanced steady-state armature currents IB = COS t 6 = * cos (t - 120) (42) ic = i cos (t + 120) are applied and the rotor is again at synchronous speed so that 0 = 00 + i (00 is the rotor position at zero time), then Ja = I COS (0 00) *6 = *cos(0 -00 - 120) (43) ic = i cos (0 - 00 + 120) and, by equations 20, id = * cos 00 iq = i sin 00 (44) i0 = 0 The armature flux linkages are fa = xdi cos 0o ^a = +xqi sin 00 (45) ^0 = 0 and the terminal voltages, with armature resistance neglected t. arc, by equations 32, ed = tq Xyi sin 00 eg = +td = Xdi cos 00 (46) e0 = 0 From equations 35, the voltage of phase a is ea = xqi sin 00 cos 0 + x,ii cos 00 sin 0 (47) Thus, if we change 60 slowly from zero to 90, ca changes from ea = -\-Xdi sin t to ea = +xqi sin t. The other two phase voltages, eb and ec, t The armature resistance is usually loss than 1 per cent, while the steady-state armature reactances are of the order of magnitude of 100 per cent.

24 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE also vary in the same way, and the flux linkages (by equations 45 and 34) as well. The direct- and quadrature-axis steady-state, or synchronous, reactances of a synchronous machine may be measured in this way by supplying the terminals of a synchronously driven and unexcited machine with balanced voltages and slowly varying the rotor phase angle 00. This is the so-called slip test, but we must note that because of the very small field resistance the slip must be practically negligible for it to be successful. Short-Circuit Test for xd Another way of measuring the direct-axis armature reactance Xd is from measurements of the steady-state armature open-circuit voltages and short-circuit currents. With fixed field excitation voltage e/d, the field current is, by equation 2, or, by equation 39, //' = Rfd All amortisseur currents are evidently equal to zero. On open circuit, the armature flux linkages are, by equations 24, X Yd = Xa y or, by equations 37, / v T Xafd Yd = .X.afdlfd = fd Rfd The armature terminal voltages are, by equations 32, ed = 0 ^afd Xafd eq = - e/d = -r efd = E (48) Tfd Kfd e0 = 0 where E is introduced for convenience and may be considered as the field excitation measured in terms of the terminal voltage that it would produce on open-circuit, normal-speed operation. Note also that E is the field excitation as normally used in the steady-state vector diagram, and even for many transient problems in those cases where the effect

POWER OUTPUT 25 H=0 Xafd r Xafd 6fd - f? E Ifd = Xd Xd Rfd Xd Xd of the amortisseur may be either neglected or considered only approximately as simply a more or less arbitrarily added damping effect. If the field excitation is given as E rather than e/d, it is evident that the first equation 36 may still be written in the form ^ = G(p)E Xd(p)id, where now G(0) = 1.0 instead of xafd/rfd. For unity terminal voltage (eq = E = 1.0) the field voltage must be e/d = Tfd/xafd. On short circuit e^ = eq = 0, and with negligible armature resistance td = tq = 0. Then equations 37 and 38 give and id = //d = ^ = - = - (49) Xd Xd Kfd Xd Xd Thus Xd may be found from the ratio of the steady-state open-circuit voltage to the steady-state short-circuit current, neglecting saturation. Actually it is found convenient to calculate Xd as the ratio of the field current Ifd = Xd/Xafd required to produce unit armature current (by equation 49) on short circuit to the field current Ifd = l/Xafd required to produce unit terminal voltage on open circuit. This ratio is directly the per-unit reactance. Zero-Sequence Reactance The zero-sequence armature reactance may be measured by impressing zero-sequence currents ia = ib = ic = i cos t (and by equations 20, id = iq = 0, and i0 = i cos t). Then, by equations 36, td = "A = 0, and \p0 = xqi cos t. Equations 32 now result in ed = eq = 0, and eo = +x0i sin t + ri cos t. Note that now armature resistance may not be negligible (although it probably will be) since the zero-sequence reactance x0 is small (3 to 10 per cent) compared to Xd or xq. Power Output The instantaneous per-unit power output of a three-phase synchronous machine is given by P = \(eaia + ebib + ecic) (50) where the factor % is introduced so that with balanced operation at unity power factor and with voltages and currents of unit magnitude the power output is unity. The power is output rather than input because of the original definition of the sign of armature current (see equation 1).

26 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE Now eliminate the phase quantities by substituting from equations 33 and 35. We obtain the power in terms of direct-, quadrature-, and zero-axis quantities as P = edid + eqiq + 2eoi0 (51) For balanced current and voltage of unit magnitude (V e2d + e2q = ~Vi2d + i2q =1.0 and eo = to = 0) and of unity power factor (iq/id eq/ed), the power is again unity. As an example we shall consider the steady-state power fed into a network of negligible impedance (an infinite bus) and with a voltage of magnitude e. Suppose the open-circuit voltage E [E = (xafd/rfd)efd as in equation 48] of the synchronous machine is ahead of the corresponding bus voltage by a constant angle 5. That is, if by equations 35 the open-circuit machine voltages are ea = E sin 0 eb = -E sin (0 - 120) (52) ec = -E sin (0 + 120) Then the system voltages are ea = e sin (0 5) = + e sin (5 0) eb = -e sin (0 - 8 - 120) = +e sin (8-0 + 120) ec = -e sin (0-8 + 120) = +e sin (8 - 6 - 120) or, expanding the sines, ea = e sin 8 cos 0 e cos 8 sin 0 eb = e sin 8 cos (0 - 120) - e cos 8 sin (0 - 120) ec = e sin 5 cos (0 + 120) - e cos 5 sin (0 + 120) whence, by comparing with equations 35, we see that ed = e sin 8 eq = e cos 8 (53) e0 = 0 In the steady state the currents id and iq may be found from equation 32: ed = -tq - rid eq = +fa nq (54)

POWER OUTPUT 27 where, by equations 37 and 38, fa = +Xafdlfd Xdid By equation 39, By equation 48, Then or /Xafd\ fa = + [ -jt- J efd - x&d fa = E Xdid fa = Xqtq eq = E x^d riq red + xq(E eq) id = to = XdXq + r2 +xded + r(E - eq) xdxq 4- r2 (55) (56) (57) In terms of the bus voltage e, as given by equations 53, the currents are re sin 5 + xq(E e cos 5) U= t0 = XdXq + r2 +Xde sin 5 + r(E e cos 5) (58) XdXq + r2 The power output is, by equation 51, P = e<{id + eqiq + 2e0i0 re2d + xq(Eed eqed) + xdedeq + r(Eeq e2q) XdXq + r2 or, by rearranging and using equations 53, e2 Ee(xq sin 5 + r cos 5) rer + (xd xq) sin 25 "2 -r P = (59) XdXq + rz The power input may be computed by adding to P the armature t2r losses. These armature losses are (see equations 33): f r(t2O + i2b + i2c) = r(i2d + i2q + 2i\)

28 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE Since the machine is operating at synchronous speed the per-unit power input is numerically equal to the per-unit torque. If armature resistance is neglected (r = 0) the power relation reduces to Ee (xd - xq)e2 P = sin S + sin 25 (60) This is of course also the steady-state torque in per unit, since the rotor speed is unity and since losses have been neglected. Torque A general expression for the torque, valid under both transient and steady-state conditions, may be derived in two ways: | first, by using the expression for power (equation 51), and, second, directly from a consideration of the forces acting on the armature conductors. The power output is P = edid + eqiq + 2e0i0 (51) By equations 32, ed, eq, and e0 may be eliminated to obtain P = id(ptd ~ tqp6 - rid) ~ riq) + 2i0(pt0 - n'0) or P = dd qq 00 qd ~ dq ~ d q (61) This equation may be interpreted as (net power output) = (rate of decrease of armature magnetic energy) + (power transferred across air gap) (armature resistance loss). From this it is evident that by dividing the air-gap power (the second term on the right side of equation 52) by rotor speed (p6), we obtain the torque T as T = Iqtd ~ i^q (62) For balanced normal-speed operation with no armature losses we may write \l/d = eq and \l/q = ed from equations 32, whence the torque is T = v, + ided (63) which checks equation 51 as it should. Incidentally, this has also demonstrated that the torque defined by equation 63 is unity for unit load and speed. J A third derivation of the torque, from energy considerations, is given in Appendix B.

TORQUE 29 By the second method we may consider that there is distributed around the air gap of the machine a flux wave having a density proportional to B, where B = fad cos 7 + faq cos (7 90) (64) = fad cos y + faq sin y where y is an angle measured from the direct axis in the direction of rotation in Fig. 5 and fad and faq are the space-fundamental components of air-gap flux, fad and faq are obtained by substituting xad and xaq for xd and xq, respectively, in equations 37 and 38 for fa and fa^. That is, they do not include the armature-leakage flux linkages xtid and xiiq (where xi = Xd Xad = xq xaq) which do not link the rotor. Review of the derivations of equations 9, 11, and 14 will show more clearly the significance of the leakage reactance. Similarly, by considering the space-fundamental distribution of armature current along the air gap we may obtain an expression for the current-density distribution as i = id sin y + i sin (7 90) (65) = id sin 7 tq cos 7 We have chosen current into the paper in Fig. 5 as positive and remembered also that positive id a,ndiq produce negative flux linkages. The torque is proportional to the total force acting on the armature current and thus to the integral from 0 to 2w of the product (Bi) f*y = 2x py = 2ir T oc I iB dy = f (id sin 7 iq cos y)(fad cos 7 + faq sin 7) cfy Jy=0 Jy = 0 T CCw(icCtao tqfad) Since positive current was chosen into the paper in Fig. 5 and positive flux density is directed radially, positive force acting on the armature is directed opposite to the direction of rotation. The torque of equation 62 is acting on the armature in the direction of rotation. Also the constant iv may be disregarded since we have been concerned only with proportionalities. In effect then we have found a formula for torque: T = iqfad - ictiaq (66) Now, since fad = fa + xtid and faq = tq + xiiq, we may write T = iq(fa + xtid) id(fa. + xiiq) = Iqfa tdfa. (62)

30 MATHEMATICAL DESCRIPTION OF A SYNCHRONOUS MACHINE We may observe further that any pair of quantities Kid, Kiq may be added to fa and il/q, respectively, without affecting the torque. In the case of a magnetically round-rotor machine, for example, in which Xd = Xq, K may be taken equal to Xd. Then the stator currents id and iq are eliminated from the flux linkages, so that they depend only on rotor currents. This corresponds to the absence of the so-called reluctance torque. Summary In this chapter we have developed a set of equations for the performance of a salient-pole synchronous machine. Relations among the speed, voltages, currents, power, and torque have been obtained. For the case of constant speed the voltage-current equations obtained are linear differential equations with constant coefficients. This has been accomplished by the introduction of substitute variables id, iq, i0, etc., in place of the original variables ia, ib, ic, etc. Moreover the d and q quantities have been given a simple physical interpretation as related to magnetomotive forces acting along two physical axes of the rotor. This has simplified the flux-current relations remarkably compared to their original form and has also led to an extremely simple formula for torque. The principal requirement the machine must fulfill in order that these equations be valid is that its armature windings be effectively sinusoidally distributed along the air gap as far as all mutual effects between rotor and stator are concerned. We have been practically forced to use the d, q, and 0 axis quantities for a salient-pole synchronous machine in order to reduce the performance equations to a manageable form. However, we have noted in passing that even in a symmetric-rotor machine a considerable simplification is effected by a similar transformation. The relations and algebraic manipulations are of course much simpler for the symmetricrotor case, and we might therefore wonder if it would not perhaps have been easier to have considered this case first instead of the more general salient-pole case. In the symmetric-rotor example we could have considered the components of armature current along any two mutually perpendicular rotor axes, rather than necessarily the pole and interpole axes. Indeed, in a wound-rotor induction motor we might have been unable to decide whether to change over the stator current to axes fixed in the rotor or to change the rotor currents to axes fixed in the stator. It is this very lack of compulsion in the simple cases that has decided us to consider immediately the most general case, where we can much more easily

PROBLEMS 31 appreciate the logic and practical necessity of each succeeding step as we come to it. We have completed the development of the theory. The rest of the book is primarily concerned with applications. In the following chapter we consider certain aspects of the steady-state performance (currents, excitation, and torque) and include a vector diagram and some numerical examples. Problems 1. Consider a wound-rotor induction motor having both a three-phase stator winding and a three-phase rotor winding, so that the rotor-voltage equations are identical in form to the stator-voltage equations (equations 1). Develop a set of differential equations with constant coefficients (for constant rotor speed p6) in terms of axes fixed in the rotor by a process similar to that followed in this chapter for the salient-pole synchronous machine, namely: (a) Write out the voltage equations for the six coils of the machine, as in equations 1. (6) Write out the six flux-linkage relations, as in equations 5. (c) Note that all six-coil self-inductances are constant and that all six mutual inductances among pairs of stator coils and among pairs of rotor coils are constant since both rotor and stator are symmetrical about any axis. (d) Write down the nine sinusoidally varying mutual inductance coefficients between each rotor coil and all the stator coils, as in equations 16, taking the rotor angle 6 as the angle of the axis of phase A of the rotor ahead of phase a of the stator in the direction of rotor rotation. (e) Change over to d-, q-, and 0-axis quantities by using the transformation equations 20, 23, and 28 for the stator currents, flux linkages, and voltages and similar transformation equations for the rotor quantities except that now 0 = 0. 2. For the same wound-rotor induction motor as in problem 1, develop a set of differential equations with constant coefficients (at constant rotor speed) in terms of axes fixed in the stator. The procedure is identical except for step e. Now, instead of step e, change over to axes which may be called a, $, and 0 axes, and which are fixed in the stator, by using (a) For the stator quantities, transformation equations similar to equations 20, 23, and 28 except that 0 = 0. (b) For the rotor quantities, transformation equations similar to equations 20, 23, and 28 except that (0) is substituted for 0. 3. Discuss the physical interpretation of the currents developed in problems 1 and 2, as in the text following equations 20. Explain from a physical standpoint the change of the sign of the angle used in the transformation equations in problem 2, item 6. 4. A two-pole, three-phase synchronous machine operates at 3600 rpm. The maximum value of the mutual inductance between the field winding and any one of the three ^-connected armature coils is 0.04 henry. What is the required field current at normal voltage (13,800 volts rms line to line) and open circuit.

3 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION The Steady-State Vector Diagram In the previous chapter we have shown (equations 58) how to calculate the components of current id, iq for a synchronous machine connected to an infinite bus, when the machine excitation E (measured in terms of the open-circuit generated voltage) and bus voltage e are known. The procedure is straightforward but somewhat cumbersome unless armature resistance r is neglected, in which case equations 58 reduce to E e cos 5 id = xd (67) e sin 6 On the other hand, the converse and more common problem of calculating the excitation E and the rotor angle S for assigned balanced terminal voltages and currents is not so straightforward because now we do not know the positions of the machine axes relative to the bus voltage and so cannot immediately resolve the current and voltage into their direct- and quadrature-axis components. For this purpose it is convenient to construct a vector diagram. Observation of equation 56, ed = Xqiq rid (56) eq = E Xdid riq shows us that it is possible to consider the voltages and currents as vectors in a plane having d and q co-ordinate axes mutually perpendicular and oriented in exactly the same way as the d and q axes of the machine itself as given by Fig. 5. 32

THE STEADY-STATE VECTOR DIAGRAM 33 If we define a voltage Eq as Eq = E - (xd - Xq)id (68) then equations 56 may be written as ed = xqiq rid (69) eq = Eg xqid riq and complex voltages and currents may be denned as e = ed + jeq (70) i = id + jiq In terms of these complex values, the two equations 69 may be combined by multiplying the equation for eq over by j and adding, whence e = jE, - (r+ jxji (71) Now it is evident that, since the transformation equations for both voltage and current are of the same form, the phase displacement between e and i is the same as that between ea and ia, etc., so that, even though we do not yet know the position of the direct axis, equation 71 shows us that we may simply add the i(r + jxq) voltage drop to the terminal voltage e and obtain a quantity Eq which, although it may have no simple physical meaning, is known to lie along the quadrature axis and is moreover simply related to the excitation E by equation 68 (see Fig. 6). For now that the location of the quadrature axis is known, i can be resolved into its id and iq components, and (xd xq)id computed and added to Eq to find E. Referring again to Fig. 6, one may see that a simple graphical construction for E is possible. That is, when jxqi is calculated, jxdi is also calculated at the same time. Then by dropping a line from the point (see Fig. 6) jEd = e + (r + jxd)i, perpendicular to the line passing through jEq and the origin, the intersection E is found. It should be noted that the voltages and currents are constants (i.e., direct voltages and currents) even though it is possible to treat them in exactly the same way as is conventionally done with the complex-number representation of alternating voltages and currents. However, since the diagram is formed just as with complex a-c quantities, any balanced external circuit may be added to the diagram in the conventional manner. In case of a simple tie line consisting only of a series r and x, this would have been obvious anyway, simply by adding the tie line r and x directly to the armature r and x, but it is not so directly obvious

34 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION in case of a general network including mutual inductance and capacitance. A further difference in concept should also be remarked. In the case of alternating voltage and currents, we are in the conventional "vector" diagram representing not the real voltages and currents, nor even the complete complex voltages and currents, but rather the coefficient of Vector diagram e; (w = angular frequency) in the complex voltages and currents; in the case of the direct- and quadrature-axis voltages and currents we are representing the complete voltage or current. Field Flux Linkage In many problems involving, e.g., transient-stability and voltage-dip calculations, the phenomena of interest are such that the steady-state equations and diagrams may be used, except that because of the very small field resistance the field flux linkage rather than the field current will tend to remain nearly constant. The exact way in which the principle is applied is shown later; for the present we shall only recognize that it will be of interest to be able to identify on the vector diagram a quantity corresponding to field flux linkage.

FIELD FLUX LINKAGE 35 Referring to equations 37, neglecting amortisseur currents we have ^d = Xdid + Xafdlfd (72) ^/d = Xafdid + Xffdlfd Eliminating the field current Ifd, we find, from the first equation 72: ,,d. *+3* (73) -*a/d ^ + (--S)] and, from the second equation: Xffd */d = A-afd The quantity, X afd Xffd is a short-circuit reactance of the armature direct-axis circuit, which would be measured at the direct-axis armature terminals with zero field resistance. We define Xd = x'd = transient reactance Xffd and Xafd V'/d = E'q = voltage back of transient reactance (a quantity proiid portional to the field flux linkage) Then equation 74 becomes E'q = id + x'aid (75) or, since eq = ^d riq then eq = E'q x'did riq (76) Equation 76 is just like the second equation 56; so the whole of the derivation of the vector diagram may be repeated simply by substituting x'd for Xd, and E'q for E, everywhere. [Note that now (x'd xq) is negative.] The complete diagram, showing both E and E'q, is given in Fig. 7.

36 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION QUADRATURE AXIS Fig. 7. Vector diagram We may also find, from the second equation 72, rv Xafd , x2afd . E'q = - frd = - id + E Xlld X.ffd or, by the definition of x'd, E'q = E - (xd - x'd)id (77) This may be seen from Fig. 7 to check the value previously found. Power Output The power output of the machine is for balanced operation: P = edid + eqiq (78) If a dot product is defined, as for vectors, as the sum of the products of the inphase components of two quantities e and i, then equation 78 is equivalent to P = (ed + jeq) (id + jiq) = e i (79) or, in another way, P = real part of [e times conjugate of i]

POWER-ANGLE CHARACTERISTICS 37 Thus it is seen that even for the power equations the d, q axis quantities e and i behave just like ordinary alternating voltages and currents. If resistance is neglected, it is further evident (equation 71) that jEq differs from e only by a voltage drop having no components in phase with i, and so the power output may also be written as P = (jEq).i (80) But, by equation 71, 1-*^ TO Substituting equation 81 in equation 80 gives /je \ Eqed Ege P = + (jEq) ( )=-.- = sin 5 (82) \Xq / Xq Xq where e is the magnitude of the terminal voltage as in equations 53. In using equation 82, it must be remembered that Eq does not in general remain constant; instead either field flux linkage (or E'q) or field excitation (or E) may be taken as constant. The only object in presenting equations 79, 80, and 82 is to show ways of computing power which may be convenient when using either vector diagrams or an a-c network analyzer. Power-Angle Characteristics The steady-state power-angle equation for zero armature resistance and with fixed excitation has been given in equation 60 as Ee /I 1 \ e2 P = sin 5 + ( --- ) - sin 25 (60) xd \z9 xd/ 2 If the field flux linkage is assumed to remain constant, we may derive a similar transient power-angle characteristic in terms of the voltage E'q. We start, as in equations 54 but with r = 0, with (83) e= As before, tq xqiq, but now we use the newly derived relation 75,

38 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION It is now evident that the derivation proceeds exactly as before except that we substitute E'q for E and x'd for Xd. The steps are ed Xqlq eq = E'q x'did iq = ed xq esin 5 xq t E\ ~H Id = i i d P = edd \ Vqlq ~ and, finally, E'q e cos 5 Vd ed(E'q eq) eqed X'd Xq E'qe /1 1 \ e2 P = ^ sin 5 + ( ) - sin 25 (84) xrd \xq x dt 2 To see the form of these relations 60 and 84, we may take as a specific example a machine having the constants: xd = 1.0 (85) Xq = 0.6 x'd = 0.3 r=0 (these may be taken roughly as typical for a water-wheel generator) and a terminal voltage e = 1.0. First we consider the case of no-load excitation and field flux initially, so i = 0. By equation 71, Eq = 1.0 By equation 68, E = 1.0 and, by the similar equation, Eq = E'q - (x'd - xq)id E'q = 1.0 (86) Thus the steady-state power-angle equation is P = 1.0 sin 8 + 0.333 sin 25 (87) The transient power-angle equation is P = 3.333 sin 5 - 0.833 sin 25 (88)

POWER-ANGLE CHARACTERISTICS 39 Next we consider the case of full-load excitation and field flux initially. As full load, we take unit current at 0.8 power factor, overexcited. [That is, the phase relation is such that (inductive) reactive power is consumed by the system, so the required generator excitation is greater than for the same current at unity power factor. Similarly, we shall define a generator power factor as underexcited when (inductive) reactive volt-amperes are supplied by the system, so that the required generator excitation is smaller than for the same current at unity power factor.] Since only relative angles are essential in the vector diagram, we may take e = 1.0 i = 0.8 - jO.6 By equation 71, (jEq) = e + jxqi = 1.0 + jO.48 + 0.36 or Eq = | 1.36 + jO.48 | = 1.442 By equation 68, E = Eq + (xd xq)id where, from Fig. 6, = 1.442 + 0.4 sin (5 + <t>) , 0.48 V 5 = tan"1 = 19.4 1.36 ,0.6 4> = tan-1 = 36.8 0.8 sin ( + <*) = 0.832 E = 1.774 Similarly, by equation 86, E'q = 1.442 - 0.3(0.832) = 1.193 Now the steady-state power-angle relation is P = 1.774 sin 5 + 0.333 sin 25 (89) and the transient power-angle relation is P = 3.977 sin 8 - 0.833 sin 25 (90) The powers for these four cases are plotted in Figs. 8a and 86. The corresponding armature currents i = V i2d + i2q (see paragraph following equation 20) are plotted in Figs. 8c and 8d. The effect of saliency

40 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION (xq < xd) on torque is to introduce a second-harmonic term, the reluctance torque, having its maximum at 5 = 45 into the steady-state power-angle curve. Thus the maximum power occurs at an angle lying between 45 and 90, rather than right at 90 as in a symmetric-rotor machine. Similarly, the power-angle curve with fixed field flux linkage has a second-harmonic term having its maximum at 5 = 135 since x'd < xq; so the transient power maximum occurs at an angle between 90 and 135. The angle at which maximum power occurs may be located by setting dP/d5 = 0. From equation 60, dP Ee (\ 1 \ = cos 5 + ( I e2 cos 25 = 0 (91) dS Xd \xq xd/ But cos 25 = 2 cos2 5 1, so equation 91 becomes xqE 1 cos2 5 H cos 5 = 0 (92) 2(xd - xq)e 2 or xqE f xqE f 1 cos 5= J +- (93) 4(zd-x,)e \L4(zd-x,)eJ 2 For usual values of the circuit parameters only the positive value of the radical, giving 5 < 90, will correspond to maximum power, the other root being greater than 1.0. The same analysis can be carried through for constant field flux, leading to (94) 4(*9 x'd)e \ L4(x, x'd as the desired root. Stability In the actual operation of a synchronous generator connected to a large power system, the power input is controlled by the prime mover, waterwheel, steam turbine, etc., and the rotor angle 5 will tend to advance until the power output, as given by the steady-state curves of Fig. 8, equals the input (minus the losses which we have so far neglected). If the prime-mover input becomes too large, or, as is much more likely, if the equivalent system impedance (which we have seen may be added directly to the armature reactance and resistance) becomes too large because of changes in the system, e.g., following the switching out of a faulted line, the maximum power output may be smaller than the input.

STABILITY 41 (a) Fig. 8. Salient-pole generator on infinite bus (see equations 85-98)

42 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION STEADY STATE 20 40 60 80 100 120 140 160 ISO ZOO 220 I 8-DEGREES CURRENT - ANGLE CHARACTERISTICS WITH NO-LOAD EXCITATION (c) STEADY STATE -40 -20 O 20 40 60 80 IOO 120 140 160 160 20O 220 8- DEGREES CURRENT . ANGLE CHARACTERISTICS WITH FULL-LOAD EXCITATION to FIG. 8 (continued)

STABILITY 43 (e) (/) Fig. 8 (continued)

44 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION In this case the generator angle will continue to advance because of the net accelerating torque, and the generator may eventually operate out of synchronism with the system as it will continue to accelerate until the prime-mover torque input is reduced by its speed governor. For stable operation the generator must be operating at an equilibrium point such that for any disturbance, such as a momentary small increase of input torque, the power output will also increase so as to tend to restore equilibrium. Thus the operating angle must be smaller than that given by equation 93, or in other words the slope of the powerangle curve (dP/dS) must be positive. A general study of power-system stability is outside the scope of this book. We shall however later on be concerned with the hunting and self-excitation of machines, so it is appropriate to point out now that the criterion of stability given above (dP/d5 < 0) is concerned only with the static stability of the system if the value of P is determined from the steady-state curve. If complete system equations are investigated, the power limit found may be either greater or smaller depending on circumstances; and, on the other hand, if the system is said to be stable after a calculation of a positive steady-state value of dP/dS, we are tacitly assuming that the system introduces no other torques, such as negative damping torques arising from terms depending on rotor speed rather than rotor angle, which might produce instability. Reactive Volt-Amperes As mentioned in the previous section, the power input (and thus approximately the power output) of a synchronous generator connected to a large power system is determined by the prime mover. Thus, varying the field excitation does not change the power output appreciably but instead causes the machine angle of advance 5 to vary so as to maintain constant power P. There are corresponding changes in armature current, and, if the system reactance is appreciable, in terminal voltage. For example, Fig. 8 shows that, with a steady-state power output of P = +0.8, changing the field excitation from E = 1.0 to E = 1.775 changes the angle from S ^ 30.5 to 5 ^ 19.5 and the armature current from i ^ 0.86 to i = 1.00. The terminal voltage remains constant at e = 1.0. This variation in current at constant power corresponds to a variation in terminal power factor and in the amount of reactive volt-amperes supplied to the system. Since we are concerned at present only with sinusoidal armature quantities, the terminal reactive power may be defined as the product of the terminal voltage and the out-of-phase component of armature current. It has previously been demonstrated

REACTIVE VOLT-AMPERES 45 that the phase displacement between voltage and current is the same for the direct- and quadrature-axis quantities as for the phase quantities, so reactive power may be defined similarly to real power (equation 79) as Q = e (ji) = (ed + jeq) (jid - *9) (95) The last form of equation 95 shows also that the reactive power may be regarded as Q = imaginary part of [e times conjugate of i] (96) Thus we may write ei* = (e.d + jeq)(id - jiq) = P+jQ (97) This shows also that VP2 + Q2 = | e I | i | (98) as should of course be true for any valid definition of reactive power. The first form of equation 95, Q = e. (ji), is believed to be the most generally useful and fundamental definition. This definition corresponds in fact to the usual way in which reactive power is measured, that is, by means of a wattmeter with either its current or voltage input shifted by 90. Note that our definition (i.e., shifting the current ahead by 90) means that an overexcited generator will tend to supply positive reactive power to the connected system. Figures 8e and 8/ show the variation of reactive volt-amperes with angle for fixed excitation and for fixed field flux as discussed previously. In these cases, since | i \ and P have already been calculated and e is unity, the magnitude of Q is most simply found from equation 98. The sign however must be determined from the relative phase positions of e and i. If i lags e in our vector diagram, Q is positive. By substituting equations 53 and 67 into the last form of equation 95, we may find an expression for reactive power Q at the infinite bus directly in terms of excitation and angle: Ee /I 1 \ e2 / 1 1 \ e2 Q = cos 5 - ( + )- + ( ) - cos 25 (99) xd \xg Xd/ 2 \xq Xd/ 2 This is in a form similar to equation 60 for the power P. It is not however practical in case of a salient-pole machine to use equations 99 and 60 directly to solve for Q as the excitation E is varied, with constant P. Instead, we must assume an angle 5, calculate E from equation 60 for a given P, and then Q from equations 99 or 95 for this angle. Equation 99 is only for calculating the reactive power at the infinite bus e,

46 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION or for the case of negligible system reactance. If the system reactance is appreciable, then terminal voltage and armature current must be computed and equation 95 used to obtain the reactive power. As another kind of application of these relations, we shall determine the maximum reactive-volt-ampere-absorbing capacity of a synchronous condenser connected to a large power system. Here, neglecting all losses, the power is zero so that 5 = 0, but, on the other hand, since the condenser must operate stably, dP/dS must be equal to or greater than zero. Consideration of equation 60 or of the form of the power-angle curves shows us that the second-harmonic, or reluctance-torque, term remains constant as excitation is varied and that E may become negative. Equation 91 may be solved for E, with S = 0, to obtain #min = I 1 -- ]e (100) \ Xq/ and, from equation 99, 2 Qmin - - - (101) Xq Note that if negative excitation could not be used the minimum Q would be only e2/Xd. Power-Angle Characteristics for Two Machines In this section we shall derive an expression for the power or torque for a salient-pole synchronous generator supplying power to a salientpole synchronous motor. This will be done not merely as an algebraic exercise, but rather to show how the interconnection of two synchronous machines may be made in terms of direct- and quadrature-axis quantities. The generator and motor will be called machines 1 and 2, respectively, and subscripts will be used accordingly. The terminal voltage relations are ebl = eb2 (102) eci = ec2 The armature current relations are *ol = ~ia2 ibi = -ibt, (103) id ~ic2 The angle relation is 01 = S + 02 (104)

POWER-ANGLE CHARACTERISTICS FOR TWO MACHINES 47 Now equations 35 and 33 are substituted into equations 102 and 103, respectively. Whence, after some manipulation, we obtain (a) (b) (c) (d) (e) (f) (g) (h) (i) (j) edi = +ed2 cos 5 + e,2 sin 5 e,i = ed2 sin S + e,2 cos S COs S cqi sin 5 sin A + e,i cos 5 idi = id2 cos 8 i,2 sin S iqi = +id2 sin 5 iq2 cos 5 id2 = idi cos 5 + i,i sin 5 1,2 = idi sin 6 iqi cos 5 (105) The relations 105 may also be derived directly from Fig. 9 since the form as well as the derivation of the ed and eq relations 28 and of the d -AXIS OF MACH I FIG. 9. Terminal-voltage relations for two machines id and iq relations 20 show that they may be regarded as projections of the magnitudes of ea, e6, ec, ia, ib, ic, on mutually perpendicular d and q axes for the two machines, if ea, e\,, etc., are plotted along three equally spaced a, b, c axes common to the two machines.

48 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION In the steady state with zero armature resistance (r = 0) and at synchronous speed (p6 = 1.0), for either machine, ed = -tq = +xqiq and (106) eq = +td = E xdid Then from equations 105a and b, we find Xqiiqi = Xq2iq2 cos 5 + (E2 Xd2id2) sin 5 El Xdiidi = Xq2iq2 sin 5 + (E2 xdtf.dt) cos 5 Substituting equations 105g and h in the last two equations, and rearranging and combining terms, we find Xd2 Xq2 . idi sm 25 A / *d2 i* Xg2 *(12 ^o2 \ + iqi { Xqi H cos 25 I = E2 sin 5 \22/ / . x (107) / a'(i2 T Xg2 Xd2 Xq2 \ + ldl [ Xdl H i COS 25 I V22/ iqi sin 25 = El E2 cos 5 from which Ei[xqi + \(xd2 + xq2) %(xd2 xg2) cos 25] E2(xqi + xq2) cos 5 (108) - Xq2) sin 25 + E2fadi + ^ sin 5 where A= - \(xdi - xql)(xd2 - xg2) cos 25 (109) The power may be found from P = ediidi + eqiiqi, or, from equation 106, P = Eliql (Xdl - Xql)idliql (110) Substituting equation 108 in equation 110, we find the final expression for power: i - xgl)(xd2 - xq2)] sin 5 + Xqi) + E22(xdi - Xqi)(xdi + xq2)] sin 25 2(xdi - xql)(xd2 - xq2) sin 35} -H A2 (111) where A is given by equation 109.

POWER-ANGLE CHARACTERISTICS FOR TWO MACHINES 49 It should be evident how expressions similar to equation 111 can be found for the condition of constant field flux linkage in either or both of the two machines. Figure 10 shows plots of steady-state and transient power, current, and terminal voltage for two equal machines, each having the constants of equation 85, Xdi = xd2 = 1.0, xqi = xq2 = 0.6, and for the two conditions of no-load excitation (Ei = E2 = 1.0) and of full-load, unitypower-factor excitation. The full-load excitation is found by taking for initial condition the terminal voltage and armature current in phase and both equal to unity. Then, from equation 71 or Fig. 6, Eq = | 1.0 + j'0.6 | = 1.166 (tx \ tan-1-J = 0.513 ^term / Ex = E2 = Eq + id(xd - xq) = 1.371 (a) Power-angle characteristics with no-load excitations (6) Power-angle characteristics with full-load excitations Fig. 10. Two salient-pole machines (see equations 85 and 106-111)

50 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION 3.5 TRANSIENT 1 CO l2 I11 DC 5 ..5 u STEADY STATE 20 40 60 80 I00 I20 I40 I60 I60 200 220 8 - DEGREES (c) Current-angle characteristics with no-load excitations 20 40 60 80 I00 I20 I40 I60 I80 200 220 8 - DECREES (d) Current-angle characteristics with full-load excitations Fig. 10 (continued)

POWER-ANGLE CHARACTERISTICS FOR TWO MACHINES 51 X -T STEADY STATE \ 0 20 40 60 60 100 120 140 160 180 200 220 8 - DEGREES Terminal-voltage-angle characteristics with no-load excitation ,^ ^ s X^ N \ *-rt.B TRANSIENT' X^ {TERMINAL VOLTAGE e \ 0OO r* * o> STEADY STATED V \ \ N \ \ / 0-20 0 20 40 60 80 100 120 140 160 180 200 22C 8 - DEGREES N \\ / (/) Terminal-voltage-angle characteristics with full-load excitations FIG. 10 (continued)

52 STEADY-STATE, BALANCED, SYNCHRONOUS OPERATION Summary In this chapter we have shown how the conventional vector diagram of a-c circuit analysis may be applied to the synchronous-machine quantities, even though the latter are d-c quantities. We have found it convenient to introduce a construction voltage Eq back of xq and have pointed out that use of this voltage may be desirable when setting cases involving salient-pole machines on the a-c network analyzer. As a specific application, and largely to show the effect of saliency as distinguished from round-rotor theory, we have computed several torqueor power-angle characteristics. These characteristics have also brought to our attention the fundamental problem of securing stable operation. It has further been observed that a synchronous machine supplies not only power but reactive volt-amperes to the system to which it is connected, and that the amount of the reactive volt-amperes is largely determined by the amount of machine field excitation. In the next chapter we shall begin the discussion of short-circuit currents. Problems 1. For a synchronous machine having the characteristics: Xd = 1.3 pu xq = 1.25 x'd = 0.17 r = 0.01 and operating under the conditions, e = terminal voltage = 1.0 111 = magnitude of armature current = 1.0 plot the per unit values of the field flux linkage E'q and of the field excitation E as functions of load power factor. These machine characteristics may be representative of a turbine generator. 2. Repeat the calculations of problem 1 for the machine characteristics: Xd = 1.0 pu xq = 0.6 x'd = 0.3 r = 0.01 These machine characteristics may be representative of a salient-pole waterwheel generator.

PROBLEMS 53 Compare the values of field flux linkage and of machine angle found in these two problems. 3. Consider the synchronous machine of problem 2 connected to an infinite bus through a reactance of xe = 0.25 p u on the machine base. The machine terminal voltage is unity, the power is 0.5 p u. With the field excitation reduced, and the infinite bus voltage correspondingly raised so as to maintain unit terminal voltage, until the machine is operating at its steady-state stability limit, calculate the field current E< the field flux linkage E'q, and the reactive power at the machine terminals. Neglect armature resistance.

4 THREE-PHASE SHORT-CIRCUIT CURRENT It is necessary to understand the nature of the short-circuit currents of synchronous machines, particularly for large machines, in order to apply proper relaying and switchgear and to evaluate the winding stresses and shaft torques incident to short circuits. In a large power system, there may be considerable interaction and swinging among the various machines following a severe system short circuit, and there may be voltage-regulator action. So we must regard the analysis now to be presented merely as a necessary beginning for a more complete study. For the study of whole systems, each machine cannot be treated in detail, but, before we can use the conventional approximations intelligently and confidently, we must know what we are missing. For this reason we consider first a three-phase short circuit at the terminals of a single synchronous machine. Consider a synchronous generator initially operating at balanced load and running at rated speed w (w = 1). The terminal voltages will be (see equations 53) Sd = ed0 = e sin 6, eg = eqQ = e cos 5, e0 = 0 To find the short-circuit currents, we apply unit-function terminal voltages equal and opposite to those existing before the short circuit. Thus equations 32 become ed0l = e sin 51 p\l/q rig + i/yw = eg0l = e cos 51 (112) pt0 n'0 = 0 The currents and flux linkages in equations 112 are only the components caused by the short circuit, and the currents found from these equations must be added to the initial load current to find the total current. Since we are concerned only with the changes of flux linkage, and since we shall assume constant excitation (or no voltage-regulator action), equations 36 become td = -xd(p)id, tq = -xq(p)iq, t0 = 0 (113) 54

THREE-PHASE SHORT-CIRCUIT CURRENT 55 It is evident from equation 112 that no zero-sequence effects are produced by simultaneous short circuit of all three phases, and so we need consider only the direct- and quadrature-axis equations. By equations 113 in 112, r]id uxq(j>)i (114) [pxq(p) + r]iq = whence . _ [pxq(p) + r]eM + [pxd(p) + r][pxq(p) + r] + <a2xd(p)xq(p) (115) - uxd(p)ed0 [pxd(p) + r][pxq(p) +r]+ u2xd(p)xq(p) The solution of these equations is straightforward but extremely complicated, as they may be of a very high degree in the differential operator p. It will therefore be of interest to examine the nature of the solution from a physical point of view and thereby to arrive at a reasonably simple formula. First, in order to show the degree of the equation, we consider a specific case in which the machine has one rotor circuit in each axis in addition to the field. Then from equations 39, with I2d = hd (etc.) = 0, and substituting for tfd and \l/id from equations 37, we find, for the field and direct-axis amortisseur currents,* Ifd = - - - - (116) T lid = - - (117) A(p) where A(p) = p2(xndxffd - x2fid) + p(xndRfd + xffdRid) + RuRfd. Substituting equations 116 and 117 in the first equation 37 for \l/d, we find P(XlidXafd XfidXald) + XafdRld efd A(p) itr^afdRld + X2aldRfd) > 77 *d (H8) A(p) * Note that in these equations, as in all the rest of this book, only the reciprocal system of inductance coefficients is used, but we have returned to the use of lower case x's, as remarked in Chapter 2.

56 THREE-PHASE SHORT-CIRCUIT CURRENT By comparing equation 118 with the first equation 36, it becomes evident that _ _ p(XlldXafd - XfidXaid) + Xf _ 2 2 P(XlldRfd (120) Similarly, for the quadrature axis, we may start from equations 40 (with I2q = Izq = 0) ar>d equation 38. The final expression for the quadrature-axis operational impedance is *9<P) = *9 -- PXo1g (121) Now returning to the operational expression 115 for the short-circuit currents id and iq, we can see, by substituting equations 120 and 121 for Xd(p) and xq(p), respectively, that the solution requires finding the roots of a fifth-degree polynomial in p. We observe further that if armature resistance is neglected (r = 0) the common denominator of equations 115 immediately reduces to (p2 + u?)xd(p)xq(p) the roots of which can be found by solving two quadratic equations and one linear equation. Neglecting armature resistance means that the fluxes linking the armature circuits cannot change from their initial values after the short circuit occurs (see equations 1) so that there is a d-c flux (and a corresponding direct current) in the armature that never decays. The actual current will be smaller than the current so found. The short-circuit currents may be written as (122) (P2 + o, To find an upper limit for the current, we may also at first neglect all rotor resistances. In this case, equations 120 and 121 lead to

THREE-PHASE SHORT-CIRCUIT CURRENT 57 Xd(p) = X"d = Zd(o) 2 Xf x2 = X"q = !,() = Xq - -i-* (124) Equations 120 and 121 show that neglecting rotor resistance is the same as taking p = , and this in turn shows that we are computing initial values, since taking p in the notation of operational calculus corresponds to taking initial values of a function. Indeed, without dragging in the mathematical notation at all, we may find it very plausible to believe that by neglecting all resistances we can compute initial values of current, since with zero resistance the fluxes linking every circuit remain constant, as they do at the first instant even with resistance.10 The actual short-circuit current may thus be expected to tend at first to follow a course given by equation 1 15 with Xd(p) = x"d, xq(p) = x"a, but gradually to deviate from this value and decay eventually to a final steady-state value obtainable by putting p = 0 in equation 115 or by using the steady-state equations directly. We define x"d = direct-axis subtransient reactance x"q = quadrature-axis subtransient reactance The name "subtransient" is used in order to distinguish these reactances from the transient reactances, which are defined in the same way except that the presence of amortisseur windings is ignored. Historically, the machine without amortisseur was analyzed first and the name "transient" appropriated for that case. From equations 122, with xd(p) = x"d and xq(p) = x"q, and with <o = 1.0, td0 . e,n id = sin t + (1 - cos 0 xdxd eg0 , d0 iq = sin t -- (1 - COS 0 XqXq or since (equations 53) ed0 = e sin 5, e90 = e cos 5 e id = -- [cos 5 cos (< + 6)] x" d (125) e iq = -~ [sin S sin (< + 5)]

58 THREE-PHASE SHORT-CIRCUIT CURRENT By equations 33, the current in phase a is ia = id cos (t +00) - iq sin (t + 00) (126) or 1 1 \e / 1 1 \ e 22 d X q/ Z \X d X q/ i X [cos (t + 00 + 5) - cos (2< + 00 + 5)] (127) The currents in phases 6 and c are obtainable by replacing 60 by (60 - 120) and (00 + 120), respectively. For no load initially, 5 = 0 and ia reduces to e/11\e ia = COS (t + 60) ~ ( + ) - COS 6Q X d ^^ d < *" cos(2'+w (128) For short circuit from no load, equation 128 shows that the initial armature current is composed of a fundamental-frequency component depending only on x"d, a d-c component depending on an average of x"q and x"d, and a small double-frequency component depending on the difference between x"d and x"q (the subtransient saliency). As the other extreme we may take the case of short circuit from a steady-state load with 5 = 90. It must be remembered however that this may be an unstable condition and is considered only to show the limiting formula for short-circuit current. For 5 = 90, equation 127 becomes 1 \e 7rjo q/ & I 1 \e --- ^)-sin(2<-r.00) (129) X d X q/ 2 The principal effect of initial load angle is seen to be to make the fundamental component of short-circuit current depend on x"q rather than on x"d as in the no-load (S = 0) case. This point is mentioned because there has been some tendency to assume that short-circuit current always depends principally on the direct-axis machine constants. Synchronizing Currents Another application of equation 127 for an even greater effective angle B may arise in synchronizing a generator. In this case let us assume that

STEADY-STATE COMPONENTS OF SHORT-CIRCUIT CURRENT 59 the machine is running at synchronous speed at an angle S0 ahead of the bus voltage. Then the machine voltages are egm = E, edm = 0 and the bus voltages are eqb = e cos 5o, edb = e sin S0 When the machine terminals are connected to the bus, the resulting currents are as before found by the application of voltages equal and opposite to those existing across the open switch. These voltages are C,O = tqm tqb = E e COS S0 ed0 = edm edb = e sin S0 Now eq0 is likely to be smaller than ed0, and ed0 may equally probably be either positive or negative; so the effective angle may be very large in either direction. Since both voltages will usually be rather small for a proper synchronization procedure the currents will also be relatively small. However, it is evident that synchronizing with an infinite bus out of phase may in the limit (at SQ = 180) be equivalent to short circuit from double voltage. On the other hand, it should be further evident that, if the bus or system to which the machine is being synchronized has a reactance equal to or greater than the direct-axis subtransient reactance of the machine, the maximum current will not exceed the three-phase short-circuit current. Steady-State Components of Short-Circuit Current In the steady state, from equation 115 with p = 0, and w = 1.0, re sin 5 + xqe cos 5 id = a~i r" + xdxq (130) re cos 8 xde sin 5 r2 + xdxq The current in phase a is e [ Xd + xq ia = -y ~r sin (t + 60 - 8) H cos (t + 60 - 8) r2 + xdxq L 2 - cos (t + 00 + 8) (131) 2J

60 THREE-PHASE SHORT-CIRCUIT CURRENT For 5 = 0, ia = -5 - [~r sin (t + 00) + xq cos (t + 00)] (132) r + xdxq For r = 0 but 6 ^ 0, e IV 1 1\ / 1 1\ I ia = - ( - + - cos (t + 60 - 5) + I --- ) cos (t + 60 + 5) 2 l\Xd Xq/ \Xd Xq/ I (133 The steady-state current is all of fundamental frequency and for zero armature resistance is obtainable by replacing Xd(,x>) by ,j(0) and xq(&>) by xq(0) in the fundamental-frequency terms of the initial-current equation 120. We should note that these formulas for steady-state current give only the component produced by the application of a terminal voltage equal and opposite to that existing before the short circuit. Actually, in the steady state this whole component is exactly canceled out by the corresponding components of load current produced by the terminal voltage, leaving only the component of load current produced by field excitation. Thus, as we should naturally expect, there is no evidence of the angle 5 in the total steady-state short-circuit current. This total current can be computed simply by taking 5 = 0 and e = E in equations 130-133, or by setting e = 0 in equations 58 of Chapter 2. For the present, however, we are interested only in the components produced by short circuit, since our object is to arrive at some idea of the course of this current as a function of time. First we note that the d-c and double-frequency components of the phase current ia disappear completely. Both of these components are derived from the fundamental-frequency components of id and iq, so it is evident that they must have the same rate of decay. Both of these components arise from the flux which is trapped in the armature circuits at the instant of short circuit. This flux gradually decays to zero and, considered as approximately constant, generates fundamental- (i.e., rotational-) frequency currents and fluxes in all the rotor circuits. It therefore seems reasonable that the decay of the d-c component may be computed by considering only the armature resistance, since the rotor resistances may be negligible in comparison to their reactances at fundamental frequency. We may refer back to equations 115 and now find the roots of the common denominator of id and iq: r)(x"qp + r) + x"dx" q

STEADY-STATE COMPONENTS OF SHORT-CIRCUIT CURRENT 61 The roots of equation 134 are The frequency is practically unity, and the time constant is o*." " ix dx q Ta3 = r(r 4. r" \ f\f d T X q) (136) and is the ratio of the average (i.e., as should be expected, the same kind of average that is effective in determining the magnitude of the d-c component according to equations 127 or 128) armature short-circuit inductance to the armature resistance. This approximation has been justified not only by comparison with actual short-circuit tests but also by the mathematical analysis of several cases including all resistances." In a similar way the decay of the transient part of the fundamentalfrequency component of phase current corresponds to the decay of flux linking the rotor circuits and may be approximated by examining the case of zero armature resistance. From equations 122 it is evident that there is a purely sinusoidal component of id and iq corresponding to the factor (p2 + 1) (i.e., u = 1.0) in the denominator and to the decaying term discussed in the previous paragraph. This component may now be disregarded and our attention turned to the terms arising from the factors Xd(p) and xq(p)< which determine the decay of the d-c components of id and iq. In the case of iq, from equation 121, , . gg ag qq xq(p) = - -r- (137) XUqP + Klq This gives, as the time constant of the decaying component of iq, X alq 2~ X alq _, 1 93 = XqRlq Rig X2alq JLrj .r v" XiIqX q The last form of equation 138 shows that the effective quadrature-axis rotor time constant for a three-phase short circuit is the time constant

62 THREE-PHASE SHORT-CIRCUIT CURRENT of the quadrature-axis amortisseur coil considered by itself, multiplied by the ratio of subtransient to synchronous quadrature-axis reactance. The second form of equation 138 shows that it may also be found directly from the short-circuit reactance viewed from the amortisseur circuit. In the case of id, Xd(p) contains the second power of p so that it would seem that a quadratic equation must be solved. However, consideration of the magnitudes involved permits a considerable simplification. That is, the amortisseur resistance is usually very high in comparison to the field winding resistance. For example, the ratio may be of the order of 50. Thus it is permissible to calculate the rapid decay of amortisseur flux on the assumption of zero field resistance, and the slow decay of field flux on the basis of infinite amortisseur resistance, that is, with the amortisseur open-circuited. On making each of these two assumptions in turn, we need only solve a first-order equation in p in each case. From equation 120, with Rid = , we find, for the direct-axis transient or field time constant, X afd Kfd With Rfd = 0, the subtransient or amortisseur time constant is, from equation 120, Xd(XndXffd da X2fid Rid X'd 040) The fundamental-frequency component of armature current arising from id thus decays in two stages, first rapidly from its initial subtransient value as determined by x"d to a transient value determined by the transient reactance x'd [calculated 'from Xd(p) with zero field resistance but without consideration of the amortisseur], and then more slowly from the transient value to the steady-state value as determined by Xd. The fact that the intermediate or transient value of id is determined by x'd may be seen from equation 75 of Chapter 3 and the accompanying argument, in which it is shown that, for the condition of constant field flux linkages instead of constant field current, x'd replaces Xd. From the mathematical point of view, equations 139 and 140 are to

STEADY-STATE COMPONENTS OF SHORT-CIRCUIT CURRENT 63 be regarded as approximations to the roots of the quadratic Xd(p), and these are in turn approximations to two of the roots of equation 115. As mentioned previously, the final justification for these approximations must come from comparative calculations such as are made in reference 11. The accompanying table (taken from this reference) shows the very close agreement between exact and approximate values of the several time constants for a particular synchronous machine with circuit parameters as given. SHORT-CIRCUIT TIME CONSTANTS Machine 1 2 3 4 Field Exact Approx. 242 242 485 485 248 242 490 485 Direct amortisseur Exact Approx. 7.94 8.125 8.0 8.125 Quadrature amortisseur Exact Approx. 8.75 8.75 8.75 8.75 Armature Exact Approx. 85.3 85.3 85.4 85.3 58.2 58.0 58.4 58.0 Machine 1 (No amortisseur) id = 1.2 p u xq = 0.8 x'd = 0.2909 T'do = 1000 radians r = 0.005 Machine 2 ame as machine 1, except that T'do = 2000 radians) Machine 3 (With amortisseur) xd - 1.2 Xq = 0.8 Xffd = 1.1 zii, = 0.8 Xnd = 1.1 Xalq = 0.6 Xafd = Xald = XfU = 1.0 rig = 0.04 rfd - 0.0011 r = 0.005 rU = 0.02 Machine 4 (Same as machine 3, except that rfd = 0.00055) Because of the relative complexity of equation 140, it is not obvious from this equation but it should be obvious from the physical reasoning

64 THREE-PHASE SHORT-CIRCUIT CURRENT leading up to it that T"d3 is simply the ratio of the short-circuit reactance viewed from the direct-axis amortisseur circuit to its resistance. Later, equivalent circuits for the direct- and quadrature-axis operational impedances will be given, and these will aid in simplifying the numerical work of calculating the various reactances required. We may define so-called open-circuit (that is, the armature is opencircuited) rotor time constants as q0 (141) Xnd whence equations 138-140 become mi qO x"q n _ mil _ * 3 1 O ~~ X T'd3=T''l0 (142) xd r", mi _ T" 1 d3 1 *) x'd Similar relations for other types of short circuit are derived later, so that the necessary time constants for any short-circuit case can be conveniently obtained from the open-circuit values. Finally, from the various time constants discussed previously, the approximate expression for the short-circuit current is - e-t/Ta>e cos (t + S) (143) X"d iq = _ 17 J_ - -) e-t/T"'> + -] e sin S + ~ rl/T>e sin (t + S) L \X q ZCq/ Xq J C q This is the current for short circuit from any load condition, but with small armature resistance. It is the component produced by the sudden disappearance of the terminal voltage e and does not include the original

STEADY-STATE COMPONENTS OF SHORT-CIRCUIT CURRENT 65 load components, which are, from equations 58 with r = 0, E e cos 6 (144) esm S '9 Adding equations 144 and 143, we find the total current, including the load current, to be E H T'/r" e cos (t + &) , (145) iq = - f ) e~'/T"^e sin & + e~'/r"e sin (t + 5) \X q 2Cq/ X g Under the assumption that T"d3 and T"g3 are much smaller than and in view of the method of derivation of these approximate results, it is evident that the fundamental component of armature current ia (which corresponds to the constant and exponentially decaying components of id and iq) has at first a value determined with all rotor resistances zero (that is, with flux linkages constant in all rotor circuits), then a value determined by neglecting the amortisseur circuits and the field resistance (that is, with field flux linkage constant), and finally the steady-state value. Similarly, the d-c component of armature current ia consists of a single exponentially decaying term determined by an average shortcircuit reactance. As a check from a somewhat different point of view on the magnitude of the transient component of short-circuit current as determined by the condition of constant field flux linkage we may compute the ratio of transient to steady-state current from equation 145 and show that it is exactly equal to the ratio of initial to final field flux linkage, as it should be. From equation 145, E ( 1 1\ h I 1 e cos 5 Id (transient) Xd \X d Xd' Id (steady state) E Xd d \e cos 8 -1) - (1450) 'd ) x'd E

66 THREE-PHASE SHORT-CIRCUIT CURRENT From the second equation 37, the initial field flux linkage under load is Xafdid (E e COS 5\ fx'd\ Xafd - ) = ( 1 Xffjifd H -- e cos 5 Xd / \Xd/ Xd Xd and the final field flux linkage is . Yfdf = Xffdlfd Xafd ( ) = I 1 Xffjlfd \Xd/ \Xd/ whence tfdi _ xafde cos 5 COS 5 = (145a) (1456) When decrements are included as in equation 145, it is not possible to reduce the expression for the phase current ia to as simple a form as equation 127. It is correspondingly difficult to determine the machine constants from a short-circuit test starting from load. However, a short-circuit test from no load forms a convenient way of determining the direct-axis machine characteristics. For this case 5 = 0 and E = e, and ia becomes * = [(4- - 4-) ,-"'"*+(4. l\x"d x'd/ \x'd x d Xd/ Xd ee-i/r03n/j_< _LJ\ 0 vj i (146) The two other phase currents *6 and ic are similar, but with 00 replaced by 00 - 120 and 00 + 120, respectively. Short-Circuit Test Even at.no load there remain difficulties of interpretation of shortcircuit oscillograms, and for this reason methods of test, measurement, and analysis have been prescribed in considerable detail by AIEE and ASA test codes.12 Reference should be made to these test codes before any such analyses are attempted. The general method however is as described briefly below. Oscillograms are taken of the short-circuit currents in all three phases, using current shunts rather than current transformers because of the

SHORT CIRCUIT WITH ARMATURE RESISTANCE 67 large d-c components that may exist. Care must be taken to assure that all three phases are short-circuited at the same instant,13 that the resistance necessarily inserted in the armature circuit for the shortcircuiting and measuring apparatus is small, and that a sufficiently long tune interval following short circuit is covered. The sustained shortcircuit current for the same field excitation is also recorded with the same oscillograph and measuring equipment as well as from a separate impedance test, as a check on the calibrations. The a-c components are found by peak-to-peak measurement of the current. It may be observed from equation 146 that the current contains a second-harmonic term. This will ordinarily be so small as not to invalidate the peak-to-peak measurement, but, if it is not, an estimate of the fundamental component must be made. The effects of saturation on the wave shape may in many cases be more significant than the second-harmonic term found by the theory of the idealized machine considered here. If the sustained current is subtracted from the a-c component and the logarithm of the resulting difference plotted against time, it will be observed that except for the first few cycles a straight line is obtained. Projecting this straight line back to the time of short-circuit determines (l/x'd 1/xd), while its slope determines T'd3. Similarly, taking the difference between the plotted curve and the straight line should result in another approximately straight line which may be used to find (1/x"d l/x'd). An actual synchronous machine will have not one but several amortisseur circuits resulting in many different time constants, so that the logarithm of this last difference will only very approximately be a straight line when plotted against time. The slope of this last straight line thus determines only an average amortisseur time constant. The d-c components of the three phases should have magnitudes proportional to cos 00, cos (60 + 120), and cos (00 120). As a practical matter the d-c component is usually determined as the midpoint of the two a-c envelope lines, disregarding the possible distorting effect of the double-frequency term. The magnitude of this effect is discussed in references 14 and 15. Short Circuit with Armature Resistance The foregoing discussion has been concerned only with cases in which the armature resistance has been relatively small. There are other important cases however in which this cannot be assumed. For example, it may be of interest to find the voltage dip following the application of an impedance load which may be of any power factor from 0 to 1.0

68 THREE-PHASE SHORT-CIRCUIT CURRENT but which will usually be greater than 0.2. In the voltage-dip calculations the d-c component of current is of little interest because it does not contribute significantly to the voltage when the resistance is small and because it disappears almost instantly when the resistance is large. As a typical example we may consider an induction motor at standstill suddenly connected to a synchronous generator. Suppose that on the generator base the motor may be considered as an impedance having a reactance of 0.3 and a resistance of 0.1. If the generator average subtransient reactance is 0.2, the armature time constant is only 5 radians, or less than one cycle of normal frequency. We shall therefore neglect the d-c component in what follows. As far as the calculation of armature current is concerned, the external impedance may be included with the armature resistance and leakage reactance. Using this concept and neglecting armature transients (i.e., the "d-c" transients caused by the armature resistance), it is evident that the armature current is as before first determined with flux maintained in all rotor circuits, next with flux maintained only in the field circuit, and finally as a steady-state value. Referring to the general equations 112, 114, and 115, we see that this corresponds to neglecting the ptd and p\l/q terms of equation 112 completely (because of the neglect of armature transients) and to computing Xd(p), xq(p) first as x"d, x"q, next as x'd, xq, and finally as Xd, xq. We consider only the application of load to a previously unloaded machine, whence ed0 = 0, and At first, x"dx"q II (147) req0 Next, ld = l d = Finally, ""- r2 + x"dx"q , x'dxq (148) + XdXq (149) xdxq

SHORT CIRCUIT WITH ARMATURE RESISTANCE 69 It remains to find the decrement factors connecting these three sets of current. We examine first the decrement from i'd, i'q to id,, iq,, which is determined entirely by the field and which can be seen from equations 148 and 149 to be the same for both id and iq. Since the amortisseur does not concern us now, we have, from equation 120, with Rid = : xd(p) = xd X2afdP X2afd)p + + Rf< and, from equation 141 and the definition of , Xd(p) = where +1 m, 1 d0 = (150) Also, xq(p) = xq Now id and iq with ptd = Ptq = 0, may be written as id = r2 + xd(p)xq(p) (151) _ r + xd(p)xq(p) They have the common denominator: xd(p)xq(p) =r = .2 /T'd0x'dp + X \ T'd0p + 1 The single root, corresponding to the effective field decrement factor, is P=T'd0 \r2 The effective field time constant T'dz is therefore = 1 dO r2 + xdxq (152) This is a more general expression than the middle equation 142, and of course reduces to equation 142 if r = 0.

70 THREE-PHASE SHORT-CIRCUIT CURRENT In the previous case of small armature resistance it was shown that the field short-circuit time constant could be regarded as the ratio of the short-circuit inductance viewed from the field winding to the fieldwinding resistance. Now it may be shown that the field time constant with an armature load may be regarded in a more general way as the ratio of an effective "load" inductance viewed from the field to the field resistance. This effective load inductance of the field may be calculated as the ratio of field flux linkage to field current with the load connected. From the second equation 37 the field flux linkage is Xafdid But, from equations 149 and 48, whence or r + xdxq XdXg) X2afdXq~ xdxq If* x'dxq xdxq (153) Still a third way of obtaining the effective field inductance is from the vector diagram as shown in Fig. 11. From this diagram and from FIG. 11. Vector diagram with resistance load

SHORT CIRCUIT WITH ARMATURE RESISTANCE 71 the previous discussion of Fig. 7, the ratio of the actual field flux linkage to the field flux linkage that would exist on open circuit is tfd tfd iqr + idx'd (154) tfd0 gr + dxd But it is also evident that rid = xqiq (155) or id XQ -=~ tT (156) Substituting equation 156 in equation 154 leads to the same ratio as is given in equations 152 and 153. If subtransient components of current are neglected, the armature currents are id = (i'd - ids)e-t/T'dz + id, iq = (', - *,.)-"*"" + V *o = [(i'd - ids) cos (t + 60) - (i'q - v) sin (t + &0)]rt/T'd, + ids cos (t + 60) iqs sin (t + 00) or finally, on substituting from equations 148 and 149, ~ eq0 xdxq/ r + xdxq cos ( t + 00 + tan"1 ] (157) \ xq/ We observe that there is no change of phase between the transient and steady-state components of current and that id and iq have the same decrement factor. To find the decrement factors of the subtransient component, we have to turn again to equation 151, let the field resistance be zero, and find two roots of the resulting quadratic equation obtained by consideration of the amortisseur resistance. Since there are two decrements, inextricably tied together, there remains the problem of assigning a proper magnitude to each component. This requires an actual operational solution and cannot therefore be regarded as a simple approximation. If we must resort to such a complicated procedure, we might as well solve equations 115 completely in the first place and not bother to construct these approximate solutions. Instead, we simply use an average value of the direct- and quadrature-axis amortisseur time constants (approach-

72 THREE-PHASE SHORT-CIRCUIT CURRENT ing the open-circuit or short-circuit values depending on the magnitude of the resistance) to determine the rate of decay of the subtransient component. This apparently high-handed procedure is justified by the short time involved, by the fact that subtransient effects may be neglected entirely in some calculations, and finally by the fact that for ordinary machines the use of a single time constant for the amortisseur is only a crude approximation to the many time constants that actually exist so that we should be inconsistent in applying an "exact" method to an approximate concept. Exact time constants for particular cases have been computed in reference 11. Field Current The change of field current has already been expressed in terms of the armature short-circuit current in equation 116. Moreover, comparison of equations 116 and 119 shows us that for the condition of no change in excitation voltage (which has been assumed for all cases studied so far) ltd = pG(p)id (158) f It should be evident that, since all currents arise from the same set of system equations, the same decrement factors must apply to both field and armature currents, d-c components in the field corresponding to a-c components in the armature, and conversely. Thus, if only an approximate expression for the d-c components of field current is desired, it is only necessary to evaluate pG(p) first for p = w, next for p = o but neglecting entirely the amortisseur, and finally for p = 0 [whence pG(p) = 0]. This procedure exactly parallels the calculation of x"d, x'd, and Xd required to obtain the a-c components of armature current. For the case of no amortisseur, G(p) = zaf* (159) + Rfd and so the initial change of field current is Mfd = [PG(p)]p=a Aid = Aid (160) It is of interest to note that, if field excitation is measured in terms of as is done on the rated-speed vector diagram discussed in Chapt Note that in this equation wo must use the definition of G(p) as given in equations 36 and 119, not the alternate form discussed following equation 48.

FIELD CURRENT 73 ter 3, the initial value of A//,; may be expressed in terms of quantities measurable from the armature circuit as 2 xafd A//d = Aid = (xd x'd) A^ (161) Xffd This expression applies regardless of the load angle before the fault and of the fault (or armature) resistance, both of these variables having already entered into the determination of Ai^. The vector diagram of Fig. 11 may be used to compute the change in field current with no amortisseur. Since the calculation is based on the assumption that the field flux linkage remains constant at its previous value (presumably unity), it is only necessary to scale up the diagram so that the voltage E'q is unity (or the value prevailing before the fault) and then measure directly the new magnitude of E (or xafdlfd). This is the average value of field current, not considering the fundamental-frequency component, since it is based on the a-c component of armature current, not considering its d-c component. For a short circuit from no load and unit voltage the transient armature current is l/x'd; so the change in field current, measured in terms of equivalent open-circuit voltage, is Xd ~ x'd A# = Xafd A//d = x'd The total initial field current for this case is therefore Xd E = Xafdlfd = Xafd A//d + 1.0 = X'd This may be of the order of ten times normal for high-speed machines having low transient reactance, and we must remember that there is in addition an a-c component having a peak value of (xd x'd)/x'd. If the machine has an amortisseur, examination of the various values of pG(p) discussed above or a physical consideration of the circuits involved shows us that the initially induced field current is reduced and then tends to approach the transient value of equation 161. That is, initially the induced rotor current is shared by the field and amortisseur, roughly in inverse proportion to their leakage reactances. With typical values of machine reactance, the subtransient induced field current may be reduced to only about one third of the total induced current. Then, if we had x.d = 1.2, x'd = 0.12, the initially induced additional d-c component of field current might be only about 4 times normal. (It is greater than 3 because the initial armature current is greater with an

74 THREE-PHASE SHORT-CIRCUIT CURRENT amortisseur.) Since it would rapidly rise to 9 times normal, this would not seem to be of much significance. However, the fundamental-frequency component is also reduced to 4 times normal, so the peak field current is reduced from (1 + 9 + 9 = 19) to (1 + 9 + 4 = 14) by the amortisfeur. Summary In this chapter we have studied in considerable detail various approximations to the three-phase short-circuit current. It may almost seem more trouble than it is worth to derive these approximate formulas. However, we shall find that in many cases only initial values are needed, in many other problems the times involved are so long that amortisseur transient effects may be neglected, in still others several other circuits of the power system to which the machine is connected must be considered. In this last case we shall find indispensable the idea (which should by now be firmly in mind) of considering the decay of the various components of short-circuit current as caused by the decay of fluxes initially trapped in the individual machine circuits, the time constant of each of which is determined only by its own resistance and by an inductance calculated as the ratio of the circuit flux linkage to the circuit current required to produce that flux linkage in consideration of the presence of all other circuits. This concept is also required in the study of unbalanced faults, where it will be found that approximations are even more in order. Finally, an important reason for developing the approximate formulas is to provide a method for analyzing shortcircuit test data and thus for determining the machine constants by test. That is, we construct a relatively simple formula each term of which not only corresponds to a measurable component of current but also is very simply related to the machine constants. It would be practically impossible to figure out the machine constants from test using the "exact" equations. From this point of view we might even regard the formulas as empirical, whence the constants are properly defined only by the method of test, and the accompanying theory seems only an elegant but unnecessary burden. However, reflection and experience show that the theory is essential, first to predetermine the constants in the design of machines, and second for their proper interpretation. The first point is obvious, the second may be illustrated by an example. Since the amortisseur really contributes many time constants, we might measure only the first few cycles of short-circuit current, separate the log-current-versus-time curve into two straight-line components and think we had determined a field-time constant, unless we had a sufficiently good idea of the magnitudes involved to realize the time interval

PROBLEMS 75 that must be considered to obtain a proper separation. On the other hand we must also realize the limitations of the theory, particularly in regard to magnetic saturation. Problems 1. A synchronous machine having the characteristics, x"d = 0.30 p u x"q = 0.40 and operating initially at no load, normal speed, and unit terminal voltage, is suddenly synchronized with a bus also at unit voltage but at an angle 50 behind the machine generated voltage. Neglecting all resistances and decrements, and assuming that 5o does not change during the time interval considered, calculate the subtransient armature current for various values of 60 from 0 to 360 for the system (i.e., bus) reactances xe = 0, 0.10, 0.20 and 0.40 p u. Calculate the armature currents both with and without the d-c component, as in some cases the armature circuit resistance may be sufficiently high to practically eliminate the d-c component. 2. A three-phase, Y-connected, 60-cycle, 50,000-kva, 13,800-volt synchronous machine has the characteristics: Self-inductance of one armature phase winding = 0.008 henry Maximum value of mutual inductance between field winding and any one armature phase winding = 0.04 henry Self-inductance of field winding = 0.24 henry Mutual inductance between any two armature windings = 0.0035 henry (a) Calculate the per unit synchronous reactance Xd and transient reactance x'd for this machine. (6) Calculate the per unit transient armature current for sudden short circuit from normal voltage at no load. (c) Calculate the per unit steady-state armature short-circuit current for a field excitation corresponding to a load of 1500 amperes rms at a power factor of 0.9 (overexcited). (d) Calculate the transient induced field current in amperes and in per unit of the no-load, normal-voltage field excitation for sudden short circuit from normal voltage at no load.

5 SINGLE-PHASE SHORT-CIRCUIT CURRENT Single-phase short circuits may be either line to line or line to neutral. The only essential difference between the two types is that the line-toneutral short circuit involves also the zero-phase-sequence impedances of the machine and of any impedance connected between neutral and ground if the fault is line to ground. They are considered together here. As in the previous chapter constant rotor speed is assumed. Line-to-Line Short Circuit For a line-to-line short circuit on phases b and c we have the conditions: eb ec = 0 (162) H + c = 0 and for no load: *o = 0 Further, if armature resistance is neglected the total flux linkages of phases b and c are maintained constant at their initial values 10 as determined by ^M and iAc0, so that tb-tc = tb0 - ^c0 (163) If the machine angle at which the short circuit occurs is 60, then v/'60 = ^d0 cos (60 - 120) - tg0 sin (6Q - 120) tc0 = td0 cos (00 + 120) - tq0 sin (00 + 120) or tb0 - tc0 = VSGfco sin 60 + ^90 cos 00) (164) Since a similar equation may be written for ^ \l/c, equation 163 may be reduced to iAd sin 6 + \l/g cos 6 = fa0 sin 60 + \l/qQ cos 60 76

LINE-TO-LINE SHORT CIRCUIT 77 Similarly, in terms of direct- and quadrature-axis quantities equations 162 reduce to ed sin 6 + eq cos 6 = 0 id cos 6 iq sin 6 = 0 (165) *0 = 0 The form of these relations suggests that it may be advisable to set up the single-phase short-circuit problem in terms of new variables to replace the direct- and quadrature-axis variables which seemed so admirably suited to the study of balanced operation of a salient-pole machine. That is, if we define new variables as ta td cos 6 \l/q sin 6 tp = td sin 6 + t cos 6 ea = <?d cos 6 ea sin 6 (166) ep = ed sin 6 + eg cos 6 ia = id cos 6 iq sin 6 ip = id sin 6 + iq cos 6 the conditions of our problem become simply CH = 0 (167) ia = 0 i0 = 0 The a, /3 quantities of equations 166 are seen to be referred to the stator or armature of the synchronous machine and may be regarded as equivalent two-phase currents which will produce the same air-gap fluxes as the actual three-phase currents. Figure 5 shows that ta is the component of flux linkage in the direction of phase a and ^ is the component in the direction of the common axis of phases b and c. The relations 167 are also seen to be those for short-circuit of the j8 phase of a two-phase machine, with a the open phase. The a-, fl-axis quantities have been defined in terms of the d-, g-axis quantities, but they may more directly be defined in terms of the original phase quantities as shown below. Either by substituting equations 20, 23, and 28 in equations 166, or directly from Fig. 5, it may easily be shown that \l/a, ^p are similar to

78 SINGLE-PHASE SHORT-CIRCUIT CURRENT td, tq but with 9 = 0. Therefore, ta = fW'a + tb COS (-120) + tc COS 120] to = ~ fW'6 sin (-120) + +c sin 120] (168) *a = !(*.- 3*6 - if.) ifo, = 1/V3 (*6 - *e) (109) Solving for the phase quantities from equation 169, by equations 34 with 6 = 0, or simply by inspection of Fig. 5, we find 1 .s/3 -^ + -^-^ + ^0 (170) A + with similar expressions for voltage and current. It has been remarked in Chapter 2 that equivalent two-phase quantities will be found much simpler to use than the original three-phase jquantities even for symmetric-rotor machines, such as induction motors.16 In this case, and with balanced stator operation, either the d-, q-, 0-axis quantities or the a-, /?-, 0-axis quantities are equally well suited. If either the rotor or stator is not symmetrical about every axis then the variables should be referred to the unsymmetrical winding in order to obtain differential equations with constant coefficients. If both stator and rotor are unsymmetrical, as in the case of single-phase faults on a salient-pole machine that we are now considering, the equations will have periodically varying coefficients regardless of the reference frame used, and an exact solution becomes rather complicated. It has appeared (equation 167) that at least the boundary conditions are very simply expressible in terms of the a, ft, 0 quantities. We therefore proceed to investigate the flux-current relations. From equations 36, since there is no change in field excitation G(p) = G(0) = Xafd/rfd. Then with (xafd/rfd)efd = E, we have il'd = E - xd(p)id tq = -xq(p)iq (171) and t0 =

LINE-TO-LINE SHORT CIRCUIT 79 By equations 166, ta = cos 0 [E xd(p)id] sin 0 [ xq(p)iq] fp = sin 0 [E - xd(p)id] + cos 0 [-xq(p)iq] or, substituting ia, ia for id, iq by means of the relations, id = ia cos 0 + iff sin 0 iq = i sin 0 + ip cos 0 we obtain \pa = [cos 6xd(p) cos 0 + sin 0 xq(p) sin 0]4 [cos 6 Xd(p) sin 0 sin 6 xq(p) cos 0]t/ 3 + E cos 0 ^,3 = [sin 0 Xd(p) cos 0 cos 0 x,(p) sin 6]ia (172) [sin 0 Xd(p) sin 0 + cos 0 x(p) cos 0]^ + E sin 0 The a, /3 components of armature flux linkage are seen to be rather complicated functions of the a, /3 components of current for a salientpole machine (only slightly less complicated than the a, b, c components). Note that the trigonometric functions cannot be simply combined because the derivative operator p acts on the products ia cos 0, etc., and we have not yet specified the exact form of Xd(p) and xq(p). Equations 172 could be manipulated by converting the trigonometric functions to complex exponential functions and applying the operational shifting formulas, but the result would be simple only for a symmetric-rotor machine. Moreover, if this conversion is made, we should also find it advantageous to convert from the a, 0 quantities to symmetrical components at the same time. We do not, however, give up hope of applying the a, 0 components. Instead, we proceed to show how an approximate solution may be obtained by following a line of reasoning similar to that undertaken in Chapter 4 for the three-phase short circuit. First, if we neglect all rotor resistances, Xd(p) = x"d and xq(p) = x"q, and the flux-current relations 172 become * = -\{x"d + x"q + (x"d - x"q) cos 26]ia - \(x"d - x"q) sin 20 in + E cos 0 (173) h = ~W'd -x"q) sin 20 4 - \\x"d + x"q - (x"d - x"q) cos 20]t> + E sin 0 In terms of a, /3, 0 quantities the voltage equations must be the same as in terms of a, b, c quantities since the relations between the a, b, c,

80 SINGLE-PHASE SHORT-CIRCUIT CURRENT and a, /3, 0 quantities are linear with constant coefficients. From the physical standpoint we say that the a, 0, 0 axes are fixed in the armature just like the a, b, c axes, so there are no generated voltages. Thus: rip (174) e0 = Pt0 - ria just as in equations 1. If armature resistances are neglected we have ^ = i/^0, and, since ia = 0, equations 173 become For the armature current: sn 6 x"d + x"q-(x"d-x"q)cos26 and for the open-phase flux linkage: <;" x"d + x"q - (x"d - x"q) cos 26 The open-phase voltage ea is given by h E cos 6 (176) at For short circuit from no load \l/q0 = 0 in equations 166, and ^00 = td0 sin 00 = e,0 sin 00 = E sin 00 so that the currents ib and ic (see equations 170 and 175) become \/3E(sin 6 - sin 00) (177) x"d + x"q - (x"d - x"q) cos 26 (178) where 6 = 6Q + t, 60 being the angle of the pole axis ahead of the axis of phase a at the instant of short circuit. It is evident that the current may be resolved into a series of odd time harmonics arising from the sin 6 term in the numerator and a series of even time harmonics arising from the constant sin 60 term (the initial trapped flux linkages) in the numerator. When 60 = 0 the initial flux linkages in the short-circuited phases b and c (or /3 in the equivalent two-phase machine) are zero, and only odd harmonics appear. When 60 = 90, the pole axis is initially lined up along the common axis of phases b and c, initial flux linkages are a maximum, and the corresponding .even-harmonic series is a maximum. The open-circuit voltage appearing across any of the

LINE-TO-NEUTRAL SHORT CIRCUIT 81 machine terminals is 90 ahead of the flux linkages, so the even-harmonic series is a maximum when the initial voltage across terminals b and c is zero. Phase Quantities All the phase quantities for a line-to-line short 'circuit may be found from equations 170 as: Since ia = t'o = 0, ia = 0 V3. tb = ta V3. ic = H Since ea = e0 = 0, 2 Since ^a = i^o, ^o = 0, eO = ea = ppa eb = \ea * = * fb 1 .s/3 = - - f<* + tao t, 1 \/3 = ~2^~^^ (179) (180) (181) We shall investigate the actual magnitudes and decrements of the shortcircuit current and open-phase voltage components in a succeeding section, but first we shall show the very similar results for a line-to-neutral short circuit. Line-to-Neutral Short Circuit For a line-to-neutral short circuit the boundary conditions are ea = 0 . . (182) ib = ic = 0 and for zero armature resistance: +a = +a0 (183)

82 SINGLE-PHASE SHORT-CIRCUIT CURRENT In terms of the a, ,3, 0 quantities already developed, we may write, for short circuit from no load, f a + 4/0 = +a0 = ^aO = ^dO COS 60 = E cOS 0O (184) and 'a = 3'a H = 0 (185) From the first equation 173 and the last equation 171 (i0 = x0if\), we now have * + *o. = -hWd + x"q + (x"d - x"q) cos 2B]ia + Ecos6 -x0io = E cos 0O or, using the current relations 185, 3"(cos 0 cos 0O) la = '(186) x" + x"q + (x"d - x"q) cos 20 + Xo K J From the second equation 173 and from equation 185, the 5 component of flux linkage is +P = ~Wd - x"q) sin 20 ia + E sin 0 or - (x"d - x",)(c&s 0 - cos 0O) sin 26 ^=x"d + x"q + (x"d-x"q)cos2e + xfi + Esine (18?) Similarly, the 0 and a components of flux linkage are Xo x0E(cos 6 cos 0O) ^o = _ xo'o = ?a = (188a) 3 x"d + x"q + (x"d - x"q) cos 20 + x0 V' and +a = -Wi + x"q + (x"d - x"q) cos 26]ia + Ecos6 x"d + x"q + (x"d - x"q) cos 26 E(cos 0 cos 0O)+.E cos 6 x"d + x"q + Xo+ (x"d - x",) cos 20 [x"d + x"q + (x"d - x"q) cos 20] cos 0O + x0E cos 0 x"d + x"q + xo + (j"d - x"q) cos 20 (188b)

LINE-TO-NEUTRAL SHORT CIRCUIT 83 The open-phase voltages may be found from the relations: e0 - ^ (189) 1 \/3 3 V3 eb = - ea H e0 + e0 = e0 H -- e0 22 22 1 \/3 3 \/3 ec = -~ea - ep + e0 = ~e0 - ep It may be seen that except for the addition of the zero-phase-sequence components the line-to-neutral case is very similar to the line-to-line case. The magnitude of the current for the line-to-neutral short circuit is V3 = 1.73 times that (equation 178) for a line-to-line short circuit if x0 = 0. If, on the other hand, x0 = x"d = x"q = x", the lineto-neutral current is proportional to E/x" while the line-to-line current is proportional to v3E/2x". Then the (l-ri)/(l-l) current ratio is 2/ Vis 1.15. The line-to-neutral fault equations may be made to look even more nearly like the line-to-line equations if we define a new pair of reactances, (x", i + x0/2) and (x"q + x0/2) for the line-to-neutral case. In fact, the first analysis " of single-phase short circuits from very nearly the present point of view used reactances of %(x"d + 0/2) and %(x"q + x0/2) for the line-to-neutral case, and of 2x"d and 2x"q for the line-to-line case. These reactances of reference 17 are what would be directly measured as the ratio of voltage to current in a test in which single-phase line-to-line or line-to-neutral voltage of rated frequency is impressed on the machine at standstill and under the present assumptions. That is, either the direct- or quadrature-axis value (or any value in between) could be found by lining up either the rotor pole axis or interpole axis with the impressed magnetic field. Moreover, the sinusoidal variation of the reactance can be directly tested by taking readings at several known rotor angles. The analysis of reference 17 used a, b, c quantities and appropriate reactances for each case. The use of the a, ft, 0 quantities makes very little difference in the complexity of the analysis but does bring out directly the relations among the effective reactances for the various types of fault. Our equations also indicate a difference in phase angle. That is, if in equation 186 (with x0 = 0) we replace (6) by (6 90), we obtain equation 178. The reason is of course only that we have chosen phase a as the reference phase in both cases. If we always took the magnetic

84 SINGLE-PHASE SHORT-CIRCUIT CURRENT axis of the short-circuited phase or phases as the reference, this difference would disappear. For the case of x"d = x"q the current magnitudes may be compared to those for a three-phase fault. We then have the relative values: xo = 0 xa = x"d 3,t> 100% 100% l-l 86.6% 86.6% Z-n 150% 100% Since the usual inherent zero-phase-sequence reactance of synchronous machines is considerably less than the subtransient reactance, it is considered good practice to ground the neutral points of any so-called solidly grounded central-station generators through a reactance large enough to make XQ x"d, so that the phase current, and thus the winding forces, for a terminal-to-ground short circuit will not exceed the three-phase value. If the neutral were directly grounded and if, for example, x0 = OAx"d (and x"d = x"q) the winding forces for a lineto-ground fault might be (3/2.4)2 = 1.56 times as great as for a threephase fault. Open-Phase Voltage for Line-to-Line Fault Equations 176 and 177 show that it is only the subtransient saliency (i.e., x"d 7^ x"g) that leads to any change in the terminal voltage of the unfaulted phase. It is further evident that we can divide both numerator and denominator of the first term of equation 176 through by x"d and thus show that E(R l)(sin 6 sin 00) sin 20 &* = + E cos 6 (190) R + 1 + (R - 1) cos 20 where -2, or that the change in open-phase flux linkage is a function only of the ratio x"q/x"d. The open-phase voltage is ea = pta = E(R - 1) {[R + 1 + (R - 1) cos 20] X [(sin 0 - sin 00)2 cos 20 + sin 20 cos 0] + 2(R - l)(sin 0 - sin 00) sin2 20} 4- [R + 1 + (R - 1) cos 20]2 - E sin 0 (191)

HARMONIC COMPONENTS OF VOLTAGE AND CURRENT 85 We may reasonably expect that maximum voltage will occur for short circuit with maximum trapped flux linkages (00 = 7r/2), since these trapped flux linkages produce additional even-harmonic terms. For this case trial of various values of time (or of 6) will show that maximum voltage occurs for 6 = 270, etc., and that ea (nuo = ea (max) = E(2R - 1) = E ( 2 -^ - 1) (192) \Xd/ If the machine under consideration is not grounded, but the fault involves ground, the voltages to ground, (ea e6) or (ea ec), will be, according to equations 180, 1J^ times as great as that given by equation 192. In the usual modern machine having an amortisseur the open-phase voltage is not very great. We have, for example: x"q/x"d ea (max) (eq. 192) 1.0 1.0 1.2 1.4 1.4 1.8 For a machine without an amortisseur these values may be much greater. The /3 component of the open-phase voltage for the line-to-neutral short-circuit case is similar in form to the a or phase-a voltage for the line-to-line case. In addition there is a zero-phase-sequence component (see equation 189) which contributes to the total phase-6 and c voltages. Harmonic Components of Voltage and Current For the line-to-line case the fault current of equation 178 may be resolved into the harmonic series (see Appendix A): . , [sin 0 - b sin 30 + b2 sin 50 - b3 sin 70 H 1 x "d + Vx"dx"q \/3E sin 00 - . [0.5 - b cos 20 + b2 cos 40 - b3 cos 60 H ] (193) where Vx"g - Vx"d b = -7== 7= (194) Vx", + Vx"d

86 SINGLE-PHASE SHORT-CIRCUIT CURRENT Similarly, for the line-to-neutral case we find 3E , -vi . ii ,, x0 X d \~ [cos 6 + b0 cos 36 + 620 cos 56 + 630 cos 76 -\ ] ZE cos 60 [0.5 + b0 cos 26 + 620 cos 46 + ] (195) where now Vx"q + 0.5x0 - Vx"d + 0.5x0 b0 = - - (196) Vx"q + 0.5x0 + Vz"d + 0.5x0 The relative magnitudes of the higher harmonic terms are somewhat smaller for the line-to-neutral case than for the line-to-line case. If we take as an example x"d = 0.25, x"q = 0.36, x0 = 0.14, then 6 = 0.091, while b0 = 0.074 = 0.816 If the machine has no amortisseur, we might have x"d = x'd = 0.36, x"q = xq = 0.64, x0 = 0.14. Then b = 0.143, while b0 = 0.125 = 0.886 The harmonic components of the open-phase voltage, for the line-toline case, may be conveniently found by working from the current harmonics, instead of by a direct analysis of equation 191. From equation 173 the open-phase flux linkage is & = - \(x"d - x"q) sin 2% + E cos 6 or, from equation 193, 1 | 2E ta= -~ (x"d ~ x"g) sin 26 - , [sin 6 - b sin 30 "."" 2E sin 00 I + b2 sin 50 ---- ] - . [0.5 - b cos 26 -j ---- ] + E cos 6 (197) Vx"dx"q I This may be reduced by well-known trigonometric relations to _ E(x d - x^ _ _ _ 2(x"d+Vx"dx"q)^ Esm60(x"d-x"q)(l-b2) +6(1 -62) cos 50 -..'] + - '- X Zi V **/ d.C q (sin 20-6 sin 40 + 62 sin 60 ---- ) + E cos 0 (198)

HARMONIC COMPONENTS OF VOLTAGE AND CURRENT 87 By differentiation a,nd combination of terms, we have the open-phase voltage: ea = ea = pta = -E(1 + 6) (sin 6 - 36 sin 36 + 5b2 sin 56 ---- ) - 4bE sin 00 (cos 26 - 2b cos 46 + 662 cos 66 ---- ) (199) The fundamental-frequency voltage in the open phase is thus increased in the ratio: 6 = T= - == = - (200) "" The harmonic form of these voltage and current equations brings out the relation between the results of the present analysis and the results of the application of the method of symmetrical components. If we define the quantity vx",ix"g as the negative-phase-sequence reactance X2 and recall from our analysis of the three-phase fault that x"d there plays the role of the positive-phase-sequence reactance xi, the fundamental-frequency component of the line-to-line current (see equation 193) becomes ib = -ic = (201) and the fundamental-frequency component of the open-phase voltage (see equations 199 and 200) becomes 2x2 ea = E (202) Xi + X2 These familiar equations 201 and 202 show that the method of symmetrical components gives the fundamental-frequency components of voltage and current correctly even for a salient-pole machine if for the line-to-line case x2 is defined as the geometric mean of the direct- and quadrature-axis reactances x"d and x"q. The factor b is also expressible in terms of symmetrical-component reactances, by multiplying both numerator and denominator of equation 194 over by V x"d, as X2 1 b = (203) x2 + xi For the line-to-neutral case the relations are not so clear-cut and a different definition of x2 is required. It is evident from equation 195 that the desired equivalence is xl ~f" X2 ~H XQ = X"d H

88 SINGLE-PHASE SHORT-CIRCUIT CURRENT or *' + ?)(*-- + 7) -T With this definition of x2, b0 is given by x2 xi Xi + X0 (205) From equations 193 and 195 it is further evident that the d-c comoonent of current is determined by the reactances: For the line-to-line case: 2x2, For the line-to-neutral case: 2x2 + XQ Referring back to the three-phase case, under the assumption that the d-c component is determined by a reactance x2, we find, from equation 128: For the three-phase case, ^- = ^(^r + ^r) (206) or OT// ~.n x2 ~ -jj-','" (207) We may go a step further and imagine a short circuit through an external reactance xe, the same in all three phases. Then in all of the three formulas, equations 208, 204, and 207, where equation 208 is (208) x"d and x"q are replaced by x"d + xe and x"q + xe, respectively, and then xe must be subtracted to obtain the effective value of x2. If the external reactance becomes very large, we have approximately, for the line-to-line case: X2 Xe = V(x"d + Xc)(x"q + Xe) 9* VX2e + Xe(x"d + X"q) x"d -*.V1+whence (209)

HARMONIC COMPONENTS OF VOLTAGE AND CURRENT 89 Similarly, for the three-phase case: 2(x"d + *.)(x", + *.) x2e + xe(x"d + x"q) Xe = x"d + x"q + 2xe x"d + x"q Xe H r r". _1_ 3." * d ~ " 9 The line-to-neutral case may also obviously be treated in the same way, or for that matter by letting x0 be large even without any external reactance. We have now four definitions of negative-phase-sequence reactance. These are compared in the following table (with x0 = Kx"d) for various values of the ratio x"q/x"dx"Jx"d Eq. 208 Eq. 204 Eq. 207 Eq. 209 1.0 1.000 1.000 1.000 1.000 1.2 1.095 1.097 1.090 1.100 1.4 1.183 1.186 1.167 1.200 1.6 1.265 1.271 1.230 1.300 2.0 1.414 1.427 1.333 1.500 Except at rather large ratios of x"q/x"d, there is seen to be little difference among the four formulas. Since many faults will be through external reactance, it is convenient to have some idea of the effect of intermediate values of xe. For example, if we take xe = 2x"d, the values of X2 calculated by the various formulas are as shown by the following table: x"q/x"d xilx"d xt/x"d xi/x"d xt/X"d Eq. 208 Eq. 204 Eq. 207 Eq. 209 1.0 1.000 1.000 1.000 1.000 1.2 1.098 1.099 1.097 1.100 1.4 1 . 194 1.194 1.187 1.200 1.6 1.286 1.287 1.273 1.300 2.0 1.464 1.467 1.429 1.500 In view of these results, for most calculations requiring the use of x2 it is sufficiently accurate to use the average of x"d and x"q as in equation 209.

90 SINGLE-PHASE SHORT-CIRCUIT CURRENT Decrement Factors The values of current and voltage calculated above are those that would exist with flux linkage maintained in every closed winding of the machine. The average flux in all circuits will actually eventually decay to zero, except in the case of the direct-axis rotor windings. The oddharmonic series corresponds to an even-harmonic series of rotor-current components; similarly the even-harmonic series corresponds to an oddharmonic series of rotor-current components. The problem of determining the actual course of the current can be approached in two ways. One is to make calculations including resistance for the symmetric-rotor case, which can be solved directly; the other is to estimate, as was done in the three-phase short-circuit case, a decrement factor for each circuit; this latter idea will be pursued first. The even-harmonic series of the armature will decay to zero approximately exponentially with a time constant depending on the ratio of the effective armature inductance to the armature resistance. The effective armature inductance for the d-c component in the line-to-line case is the ratio of the line-to-line voltage V 3E to the current, or xa = 2vx"dx"q = 2x2. The armature resistance is 2r, so the armature time constant is - (210) For the line-to-neutral case, the effective resistance may include any resistance rg of a grounding reactor. If the total zero-phase-sequence resistance is r0, this grounding-reactor resistance will be (TQ r)/3 = rg, or, conversely, the zero-phase-sequence resistance will be TQ = r + 3rg. The effective armature circuit resistance is r + rg and the effective reactance is (see equation 195) d + 0.5x0) (z"9 + 0.5x0) since the circuit voltage is E in this case, so the time constant is d + 0.5x0)(x"9 + 0.5x0) a(l-n) = - ; or " 2V(x 2r + r0 ~" 2X2 + * (213) 2r + r0

DECREMENT FACTORS 91 The odd-harmonic series will not decay to zero but will decay to a steady-state value corresponding to the original field current which existed before the fault. There will be induced in the field even in the steady state a series of even current harmonics, but the average value (i.e., the d-c component) must be the same as the original current e/d/r/d. It should therefore be possible to compute the relative values of subtransient, transient, and steady-state armature current by comparing the average values of the initial rotor currents, the transient rotor current, and the steady-state rotor current. More specifically, we must calculate the rotor mmf 's (o/,#/d + Xaidiid) and (o19*19) rather than the currents, since it is the mmf that determines the fundamental value of the armature current. At the first instant, flux is maintained in all rotor windings, and so we may find from the second and third of equations 37 that the change of field and amortisseur current in the direct axis is given by Xffdifd + Xfidild = Xafdid (214) + Xndlid = Xal or, solving for ifd and iid and combining, gives . Xafdlfd + Xaitfld = - : - o - T - ' - Id X fid) = (Xd - x"d)id (215) This may be found directly from equations 116, 117, and 120 (or 123) by neglecting the resistances. In the quadrature axis, the second equation 38, together with equation 124, gives for the change of rotor current: or 2 q = (xg - x" q)iq (216) It is now necessary to find the direct- and quadrature-axis currents id and iq in terms of the currents ia and ip. The relations are id = ia cos 6 + is sin 6 (217) ^q = la Sin 6 + Ift COS 6 Since we are to be concerned only with the average (or direct) components of id and iq, it will be best to use the harmonic form of the current equations. From equation 193, the fundamental components of current (higher odd harmonics will not contribute to the average id

92 SINGLE-PHASE SHORT-CIRCUIT CURRENT or iq) in the line-to-line case are .. sing, ia(fund) = 0 (218) un . x d + V x dX whence, by equations 217, (219) q (d-c) = From equation 215, -" (220) The steady-state value of (xafdifd + Xaidiid) before the fault is x^difd = E, so that the total average rotor mmf is all in the direct axis and is equal to ( - x"d)E ^'d+V^d^q '" \x"d Thus the rotor mmf is initially increased in the ratio, / Xd + Vx"dX"g \ \x"d + Vx"dx"J so the odd-harmonic components of armature current must eventually decay to a steady-state value obtained by decreasing the initial value in the inverse ratio. Since the initial subtransient value equals the voltage V3 divided by (x"d + V x"dx"g), the final steady-state value simply equals the same voltage divided by xd + vx"dx"q. From the symmetrical-component point of view the positive-phase-sequence reactance has increased from x"d to Xd, but the negative-phase-sequence reactance has not changed. This seems reasonable since the negativephase-sequence reactance corresponds to rapidly pulsating components of rotor current even in the steady state. In the steady state the average field flux linkage is where ifd and id are now the steady-state currents. Substituting in the steady-state value of id, we find XafdE tfd = Xffdifd --' / Xd + V X dX q

DECREMENT FACTORS 93 Clearing fractions and substituting E = xafjifd gives and since tfd = X2afd xJfd = Xd X'd (x'd + Vx"dx"q\ . r/d -1 - / ) Xffdifd \Xd + Vx'dX'q/ (221) The final average field flux linkage is seen to be reduced from its initial value Xffdifd in the ratio: (222) It is therefore evident that in the transient state, after the average amortisseur current has disappeared but before the field flux linkage has been appreciably affected, both the field and the armature current must be greater than their steady-state values in the inverse ratio, so the transient current equals the voltage V 3E divided by (x'd + W'd:c"9). The odd-harmonic components of the armature current thus behave very much like the fundamental-frequency component of armature current in the three-phase short-circuit case, at least from the present somewhat simplified point of view. As far as the positive-phase-sequence, fundamental-frequency components are concerned, it is more or less like a three-phase short circuit through an external reactance x2 = vx"dx"q, just as in the usual symmetrical-component treatment. The field time constant for the decay from transient to steady state may be seen either from equation 221, which shows that the effective field inductance is reduced from the open-circuit value in the ratio 222, or from the concept of an equivalent three-phase fault through an external reactance, to be (223) A simple extension of this reasoning shows that the decay from the subtransient to the transient value is given by an amortisseur time constant, (TV/ *i d (l-l) rri/t _. dO' (224) dX'

94 SINGLE-PHASE SHORT-CIRCUIT CURRENT The final expression for the current becomes --aV34(^-P7T^),-"''M + ( , L_W.+_L_| \rd + X2 Xd + X2f Xd + X2J X [sin 0 b sin 30 + b2 sin 50 ] y/3E sin 0O J"2 where e-'IT. (W) [0.5 - b cos 20 + 62 cos 40 ] (225) X2 = Vi'V, (208) A similar analysis can be carried through for the line-to-neutral fault. Since the steps are so nearly the same, they will only be outlined. The d-c component and all even harmonics have already been shown to decay to zero with a time constant given by equations 211-213. In connection with the odd-harmonic series, equations 214-217, defining the relations between the rotor and stator mmf and between id, iq, and ia, ip are of course independent of the type of short circuit. However the fundamental component of armature current is now (see equations 195, 185, and 204): 2E ia = cos 0, iff = 0 (226) X"d + X2 + Xq whence E id (d-c) = ; ; , iq (d-c) = 0 (227) x d + x2 + Xo The argument from here on is obviously identical with that for the lineto-line case, the final result being 1 ia = 3E l\x"d + x d + x2 + x0 X'd + x2 + x0/ + ( 1 ) f-T'i\L + \x'd + X2 + X0 Xd + X2 + Xq/ Xd + X2 + Xo X [cos 6 + b0 cos 36 + b20 cos 56 -\ ] 3E cos 60 - [0.5 + b0 cos 26 + b20 cos 46 + ] e~tlT ((-> (228) x2 + 0.5x0

FIELD CURRENT 95 where z" X X 1 d(l-n) - 1 d X d + X2 + XQ (229) J d (l-n) 1 dO and x2 is given by equation 204. Field Current The average component of the field current has already been discussed in connection with the determination of suitable decrement factors. At the first instant it is obtainable from equation 214, but more generally we may use the relation 158 (see footnote to equation 158). A//d = pG(p)id (158) If resistances are neglected, this becomes (see equation 1 19 or 214) (230) XlldXffd X fid Neglecting field resistance but also neglecting the amortisseur (i.e., Rfd = 0 and Rid = in equation 119), we find Xafd Xd X d id = - - - id (231) xffd Now it is evident from the physical nature of the problem that, if we calculate id (by equation 217) from ip (for the line-to-line fault) as given by equation 225 [ip = (2/ V S)ib], then for all harmonic terms of id the corresponding harmonic term of //d may be calculated approximately by equation 230. That is, the rotor current produced is shared by the field and the amortisseur as determined by their reactances. On the other hand, the average term of the field current must be calculated by equation 230 (and the average subtransient id) only at the first instant, and by equation 231 (and the average transient id) for the transient component.

96 SINGLE-PHASE SHORT-CIRCUIT CURRENT To illustrate the magnitudes involved, we may take as an example the following per unit machine parameters: xiu = l-l5 xJ/d = l.20 Xafd = Xa\d = XfU = l.00 Xd = l.20 x'd = 0.367 x"d = 0.279 x"q = 0.367 x2 = WW', = 0.320 Then equation 230 becomes A//d = xafd Alfd = 0.395i'd (232) and equation 231 becomes A/'/d = z*fd M'sd = 0.833t'<i (233) The average subtransient, transient, and steady-state values of id are, by equation 219 and the argument following it, and with the initial XafdIfd = E = 1.0, *'W> =-7r~l 1670 X"d + x2 1 i'd(d^ =- = 1.456 X d + X? 1 id(d-e) 0.658 xd + x2 Thus the initial or subtransient change of average field current is M"sd = 0.395i"d = 0.660 and the transient change in average field current is AI'fd = 0.833i'd = 1.213 The total transient average field current is I'fd = M'f + 1.0 = 2.213 This checks the ratio, VJ=Xdx1 = 1.520 ^^ Ifd x'd + x2 0.687 which was derived (equation 221) from a calculation of the steady-state field flux linkage.

SUMMARY 97 The average field current thus rises abruptly at first from 1.00 to 1.66, then increases further with the subtransient time constant to 2.21, and finally decreases with the transient time constant to 1.00 again. All the harmonic components of field current may be computed in the same way. We shall not derive an explicit formula for field current, since the harmonic form of id may be computed by equation 217 very simply for any particular case. Similar considerations apply to the calculation of field current for the line-to-ground fault. Summary In this chapter we have computed the short-circuit current of a salientpole synchronous machine for the two similar cases of line-to-line and line-to-ground faults through small external resistance. For these cases it has been found convenient to introduce substitute variables ia, ip, i0 to replace the phase quantities ia, ib, ic. The saving in labor is not very great, but the method has the virtue of enabling us to use the same machine constants previously defined for the threephase case. The ia, ip, variables may be regarded as currents flowing in an equivalent two-phase machine. They will be found useful in other problems. The results obtained have been shown to parallel in every respect those obtainable by the method of symmetrical components and in addition to give values for the harmonic currents and voltages. The results also provide definitions for the negative-phase-sequence reactance x2 valid in each case, showing that, strictly speaking, x2 is not simply a constant for salient-pole machines. .The decrement factors have been shown to be obtainable from the point of view of an equivalent balanced fault through a proper external reactance, since it is only the positive-phase-sequence component of the fundamental component of armature current that reacts on the average components of rotor flux.. In other words, it is only the positive-phasesequence, fundamental-frequency current that produces an effective armature reaction, all other components producing fluxes that alternate rapidly in every rotor winding. Incidentally, we may with profit now return to the three-phase short-circuit case of Chapter 4 (from no load only) and rederive the sustained and transient currents from the subtransient value by the method of the present chapter, i.e., from consideration of the rotor average mmf's. The harmonic open-phase voltages produced are not in themselves large for the usual parameters of machines having amortisseurs. They are of interest, however, because cases of high voltage have arisen in

98 SINGLE-PHASE SHORT-CIRCUIT CURRENT connection with machines not equipped with amortisseurs, when the open phase has been connected to a transmission line. The transmissionline capacitance may increase the voltage considerably as it may be sufficiently great to approach resonance with the third or fifth harmonic.18 The even harmonics do not concern us so much here because they are not sustained. Finally, we have outlined a method for computing the field current. Problems 1. Calculate the induced field current for a line-to-ground fault for the machine discussed in connection with equations 232 and 233 for various values of zero-sequence reactance XD. 2. Calculate the transient and sustained values of armature current for a three-phase fault from the viewpoint of this chapter, that is, by a consideration of the rotor mmf's (see equations 214-222). 3. Calculate and plot the curves of the armature and field currents for a line-to-line fault on a synchronous machine having the characteristics: Xd = 1.25 p u xq = 0.75 Xafd =1.10 Xffd = 1.20 No amortisseur and operating at no load and normal voltage. The short circuit occurs at the instant when the flux linkages in the short-circuited coils are a maximum. Neglect all resistances. 4. Calculate and plot the voltage on the open armature coil for the machine of problem 3. 6. The machine of problem 3 has a field current of 350 amperes at no load and normal voltage. Plot the field current in amperes for the line-to-line short circuit. 6. Consider a symmetric-rotor machine having the characteristics: Xd = Xq = 1.25 p U Xafd = Xafq =1.10 Xffd = xffq = 1.20 rid = rf<1 = 0.001 r = 0.01 For a line-to-line short circuit as in problem 3, calculate and plot the armature and direct-axis field currents neglecting all resistances. Compare these currents to those of problem 3 to show the effect of saliency.

PROBLEMS 99 7. Include the effect of the stator and rotor resistances in producing decrements, for the machine of problem 6. 8. Consider the salient-pole synchronous machine having characteristics: Xd = 1.0 Xalq = 0.4 Xffd =1.0 rid = 0.02 xnd = 0.95 rlg = 0.04 Xafd = Xaid = Xfid = 0.8 rf = 0.0004 xq = 0.6 r = 0.005 xu, = 0.7 Calculate and plot the armature current for a line-to-line short circuit at both minimum and maximum initial flux linkages.

O DOUBLE-LINE-TO-GROUND SHORT CIRCUIT AND SEQUENTIAL FAULTS In addition to the three-phase, line-to-line, and line-to-ground short circuits considered previously, a synchronous machine may have two terminals simultaneously short-circuited to ground or neutral or may have two or all of its terminals short-circuited in various sequences. We shall consider the double-line-to-neutral short circuit from no load in this chapter and shall only very briefly discuss the consequences of sequential faults and of initial load. The terminal conditions for a two-line-to-ground fault from no load on phases b and c are eb = ec = 0 (234) ia = 0 and for zero armature resistance ^b = iw, "Ac = ico (235) In terms of the a, 0, 0 quantities defined in Chapter 5 these terminal conditions become ea = 2e0 ep = 0 to = ia (236) to 2^0 = taO 2"Aoo = iad = E COS 60 is = f 00 = E sin 0O The initial or subtransient currents may be found directly from equations 173 (with "Ao = x0io = -\-x0ia) as 100

DOUBLE-LINE-TO-GROUND SHORT CIRCUIT 101 {[x"d + x"q + (x"d - x"q) cos 20] + 4x0}* + (x"d - x"q) sin 2Bif = 2'(cos 0 - cos 0O) (237) (x"d - x"q) sin 261 ia + [x"d + x"q - {x"d - x"q) cos 26]ip = 2E(sin 6 - sin 0O) whence the currents are ia = io = "{(cos 0 cos 60)[x"d + x"q (x"d x"q) cos 26] - (sin 6 - sin 60)(x"d - x"q) sin 26} h- D (238) i0 = E{(sin 6 - sin 60)[x"d + x"q + 4x0 + (x"d - x"q) cos 26] - (cos 5 - cos 9o)(x",j - x"q) sin 26) ^ D (239) where > = 2[x0(x"d + x"q) + x"dx"q - x0(x"d - x"q) cos 26] or D = x"d(x"q + 2x0) + x"q(x"d + 2x0) - [x"d(x"q + 2x0) - x"q(x"d + 2x0)] cos 26 (240) These equations may be reduced to E[2x"q cos 6 - (x"d + x"q) cos 0O + (x"d - x"q) cos (20 - 60)] 4= (241) f #[(2x" + 4x0) sin 6 - (x"d + x"q + ix0) sin 60] i + (x"d - x"q) sin (26 - 60)] I 20 = {Z4:l) It is evident that these two currents may each be resolved into even and odd Fourier series just as in the case of the single-phase faults. Moreover the second of the pair of equations 240 for the denominator D has been made similar in form to the denominator of equations 175 or 178 so that we may recognize without further consideration (at least for the odd-harmonic series), the value of the geometric ratio b, where Vx"q(x"d + 2x0) - Vx"d(x"q + 2x0) ~ Vx"q(x"d + 2x0) + Vx"d(x"q + 2x0) and we may calculate the magnitude of the various harmonics if this becomes necessary. It does not seem worth while to write down the harmonic expressions at length here as it is merely an algebraic exercise (these expressions are given in reference 19 and in Appendix A), but

102 DOUBLE-LINE-TO-GROUND SHORT CIRCUIT we shall look in some detail at the fundamental and d-c components in order to determine the decrement factors. The fundamental components are (see Appendix A) 2E cos 6 Mfund> = x"d + 2*0 + V(x"d + 2x0)(x"q + 2x0)x"d/x"q 2E sin 6 un x"d + Vx"dx"q(x"d + 2xQ)/(x"q + 2x0 There seems little doubt that we are about to discover another definition of negative-phase-sequence reactance x2. The most convenient relation from the theory of symmetrical components is that the positivephase-sequence current ii is given by h = - - - (245) X2 + X0 The positive-phase-sequence component corresponds to the average (d-c) components of id and iq, since these currents are constant with respect to the rotor and so are rotating in positive phase sequence and with fundamental frequency with respect to the stator. We refer to equations 217 and compute E "x"d + 2x0 + V(x"d + 2x0)(x"q + 2x0)x"dWfq E x"d+Vx"dx"q(x"d + 2x0)/(x"q + 2x0) (246) *9 (d-c) = 0 The coefficient of ei in equation 245 should equal the coefficient of E in equation 246. The reason for this quantitative agreement is shown later in this chapter in the section on symmetrical components. It does not seem possible to arrive at a simple formula for x2 from this equivalence, but we can calculate x2 for various values of the parameters and compare it to the values found by the previous formulas. From equation 245, solving for x2, we find x2 = (247) xi + x0 X where eE X = - = (248) ii id (d-c)

SYMMETRICAL COMPONENTS 103 Since we are interested only in relative values we shall assign all quantities in per unit of x"d and consider the five cases: x"d = xi = 1.0 x' q = 1.4 x0 = 0.01, 0.3, 1.0, 2.0, and oo For these cases: X = 1.0099, 1.239, 1.54, 1.743, and 2.183 and x2 = 1.167, 1.172, 1.173, 1.181, and 1.183 All of these values of x2 are in line with our previous results. The last value, for x0 = , of course checks the line-to-line case, x2 = V x"dx"q, while the first value, for x0 = 0, checks the three-phase case, x2 = 2x"dx"q/(x"d + x"q). From the symmetrical-component relations, we should also have (page 87 of reference 20): i2 Xq io x2 - = and - = (249) ti x2 + x0 ii x2 + x0 where i2 = the negative-phase-sequence component of current, and this may result in still different values of x2. In order to test this we must find i2, so we now proceed to determine the quantitative relations between the two-phase components and the conventional symmetrical components. Symmetrical Components Symmetrical components for three-phase systems are usually defined from a physical viewpoint as three sets of quantities (e.g., currents): the first a balanced three-phase set of normal phase rotation (the positive-phase-sequence current), the second a balanced three-phase set of opposite phase rotation (the negative-phase-sequence current), the third a set of three equal currents (the zero-phase-sequence current). It is shown that it is always possible to compose any set of three unbalanced, normal-frequency currents (or voltages) by a proper addition of the three sets of symmetrical components. If we define these components as: for the positive-phase-sequence set, ia\, ib\, ici', for the negativephase-sequence set, ia2, ib2, ic2', for the zero-phase-sequence set, iao, ibo, ico, then the three phase currents are ia = ial + ia2 + la0 ib = h\ + h2 + ibo (250) ic = ici + ic2 + ico

104 DOUBLE-LINE-TO-GROUND SHORT CIRCUIT The virtues of using symmetrical components are that in terms of these components the impedances of the usual three-phase systems either are known or can be more or less easily computed, and that well-known formulas are available to handle various cases of unbalanced operation, such as faults. In particular, if the rotor is assumed to be completely symmetrical, impedances to positive-, negative-, and zero-phase-sequence currents may be easily determined for a synchronous machine. Positivephase-sequence currents combine to produce only a flux wave rotating positively at synchronous speed. As far as the rotor is concerned, it is a constant flux and so produces no voltage drop. Thus the machine positive-phase-sequence reactance is the synchronous reactance as described in Chapter 2. In effect, the conditions are those for the "slip test" discussed there. Similarly, for a symmetric-rotor machine negative-phase-sequence currents combine to produce only a flux wave rotating negatively at synchronous speed. As far as the rotor is concerned, this is a double-frequency flux since it is rotating backward at twice normal speed. With usual rotor resistances the rotor flux linkages produced are very small, and so practically speaking the impedance met is the short-circuit, or subtransient, reactance. Finally, the zero-phasesequence currents combine to produce no flux linking the rotor, so that, regardless of rotor conditions, the zero-phase-sequence impedance is encountered. If we are to compare the two-phase and symmetrical components we must have a mathematical definition of symmetrical components. A definition in line with the foregoing physical concept, and one that is very widely used, is + aib + a2U + a2i6 + aie) (251) where (a) ia, ib, ic are complex numbers defined so that , ia = real part of ia ib = real part of ib (252) ic real part of ic and so that they have as a factor e"**, where w is the angular frequency of the quantities ia, ib, ic. For example, if ia I cos (ut + a), then ia = / cos (ut + a) + jl sin (ut + a) = TV V"'

SYMMETRICAL COMPONENTS 105 (b) a = e^/3 - - - j - (253) and finally the symmetrical components are (c) iai = real part of ii ibi = real part of a ii id = real part of ai1 ia2 ~ real Part of *2 (254) ib2 = real part of ai2 ic2 = real part of a2i2 ia0 = ib0 = ic0 = real Part of i0 Note that this is not the only possible mathematical definition, but it is believed to be the simplest, especially in view of the fact that the use of equivalent complex voltages and currents is practically universal among electrical engineers. It is customary in many calculations to leave out the factor e*ut and use merely its complex coefficients as the currents or voltages. It is important, however, in the present case to consider this factor because we want to compare our results with the instantaneous values found by the two-phase component method. Also, in many cases the quantities ii, i2, i0 (or the coefficients of r"* in these quantities) are themselves called the symmetrical components. We shall follow this practice in the subsequent development. Equations 251 may be put in the form: 1 I" 1 . . \/3 . .1 11 ~ 3 L1" 2 16 lc J 2 lb lc \ 1 f 1 V3 I 12 = - ia - - (it + ic) -j' (it - U o L 2i & \ (255) or, in view of equations 169: 12 =

106 DOUBLE-LINE-TO-GROUND SHORT CIRCUIT where, since the relations between ia, ib, ic and ia, ip, i0 are all real, the quantities ia and 10 are such that ia = real part of ia (257) ip = real part of 10 Also from equations 256: ia = ii + i2 (258) It may superficially seem that we have gone a long way around here, since if without bothering to introduce completed equivalent complex phase currents we had defined quantities ii, i2, i0 by equations 251 in terms of the real currents ia, ib, * we should have arrived at exactly equations 256 and 258 in terms of the real ia, ip, and of ii, i2. We should have been getting nowhere however since the quantities ii, i2 (although in general complex) are not the conventional symmetrical components and do not have their desirable properties of corresponding to positively or negatively rotating flux waves in electric machines. For example, if the real currents are ia = I COS ut ib = I cos (ut - 120) (259) ie = I cos (ut + 120) which are known to produce only a positively rotating flux wave; then ia = /e*" ib = /-^/V" = as/e*" (260) ie = /e+^/V6" = ale>at and, by equations 251, 11 = /e;'*" 12 = 0 (261) io = 0 On the other hand, if the real currents are used directly in equations 251 we find 0 = 0

SYMMETRICAL COMPONENTS 107 so that there is no correlation between the direction of the physical rotation and the correspondingly labeled current component. A further legitimate question is how to complete the phase currents to form the complex quantities when they are not simply sinusoidal. In most cases, however, if the currents are not sinusoidal they will have arisen as the solution of some differential equation having originally either a sinusoidal or constant forcing function, so that they contain sinusoidal factors which may be simply converted to exponentials. In the present case this question does not arise. Equations 256 have now shown us the way to go from the two-phase currents to the symmetrical-component currents. Referring to equations 243 and 244 we see that, if we let (l) = X"d Then ia = A2E cos 6 (202) ip = B2E sin 6 By equations 257, ia = A2EJO (263) ip = -jB2Eeje By equations 256, ii = (A + B)Eeje (264) ia = (A Since E(j e = C1, it is evident that (A + B) is equivalent to 1/X, as stated after equations 246, which may be rewritten in the present notation as: id (d^ = (A + E)E. Similarly: i2 XQ A B - = -- = - (265) i! x2 + x0 A + B Since (I/A) and (1/JB) are simpler than A and B, we express x2 as (266a) - \A B)

108 DOUBLE-LINE-TO-GROUND SHORT CIRCUIT or [x"d + Vx"dx"q(x"d + 2x0)/(x"q + 2x0)]2x0 , =. (260 b) f -4- \/fv" 4- 9r \<T" .4- 9r W'j/V t'O l * \^ d i ^S^O/V*** 7 I *0^ ,j/* - VVV,)(z"d + 2zb)/(z", + 2x0) It can easily be shown that the value of x2 obtained from equation 266 also lies between x2 = V x"dx"q for XQ = no and x2 = 2x"dx"q -=(z"d + x"q) for x0 = 0, just like the x2 found from equation 247. Rotor Decrement Factors Since we have shown in connection with the line-to-line fault that, from the standpoint of armature reaction on the rotor average fluxes, the machine behaves practically as if it were subjected to a balanced short circuit through an appropriate external reactance, we may write down at once the rotor subtransient time constant as T" -L T rrin rriff e /Of7\ i d(ll-n) i dQ (/Of,) X d I Xe where, for the double-line-to-ground fault, X2XQ xe = (268) X2 + XQ and the field, or rotor transient, time constant as T' 4- T rt rrtf /Oco\ d (ii-n) i d0 (/oy; Xd + Xe The transient and steady-state components of armature current may be found as in the previous chapter from a consideration of the initial rotor average mmf and the final field flux linkage. From equation 215, and adding in the steady-state value of a/dVej (or E), the total initial average rotor mmf is + Xaldild = (xd x"d)id (d-c) + E (xd - x"d)E x"d + 2x0 + V(x"d + 2x0)(x"q + 2x0)x"d/x"q f x"d + Vx77^^", + 2x0)/(x"q + 2*0) +''' (270)

STATOR OR ARMATURE DECREMENT FACTORS 109 This could be evaluated for any particular case, but it seems more advisable for a general formula to return to the symmetrical-component point of view, whence it is evident that Xafdif + Xauild Xd + Xe - = (271) E x"d + xe Similarly, the steady-state average field flux linkage may be computed as JO $ I ^g i/]d = Xffdifd (272) %d ~\~ Xq just as was done for the line-to-line case in equation 221. In both these equations x2 lies between the limits, 2x"dx"q r - ~ <x2< Vx"dx"q (273) XdIXq Thus the complete course of the odd harmonics has been determined. They may be written as 2x"qEcos6 \(a x"d + xM , T ,J5/coser/ x"d + xe\ x'd + x, X"d + Xe (X 'd + Xe X d + Xe\ \ x'd + xe xd + xe / (274) xd + xe where XoXn xe = L2- (268) x2 + x0 as before, and the time constants are given by equations 267 and 269. Similarly: 2(x"q + 2x0)E sin 6 *0(odd) = D I x"d + xe\ \ X'd + xe / t/T" . (X d v Xe X d\ Xe\ ^,/rp, + I 1 e (/i " <-) + \ X d + xe Xd + xe / %d i *e (275) Stator or Armature Decrement Factors To compute the armature time constants it is convenient to think of the two components of current ia (or i0) and ip as flowing more or less independently in the armature circuits. It at once becomes evident

110 DOUBLE-LINE-TO-GROUND SHORT CIRCUIT that the a component corresponds to a current flowing through two armature circuits (phases b and c) in parallel and through the neutral. The reactance of this path per circuit is 2 x2 + 2x0 x2 + ~ O0 - *) = OO On the same basis, the resistance of this path per circuit is 2 r + 2r0 r + 2re = r + - (r0 - r) = OO Thus the armature time constant for both ia and i0 is X2 + 2x0 Taa (U-n) = TaQ (_) = (276) r + 2r0 The ft component should be similar to that for the line-to-line fault, and so the time constant should be 2x2 x2 Tap (u-n) = = (277) 2r r Two facts should be noted about these armature time constants. First, they may be widely different, as there may be many times as much resistance in the ground connection as in the armature winding, so there is no such thing as a single characteristic armature time constant for a double-line-to-ground fault. Second, we have still to determine an appropriate value of x2. We must therefore calculate the effective reactances to the d-c components of ia and ip from equations 241 and 242, instead of merely writing down expressions in terms of an undefined x2. From equation 241, by integration the d-c .component of ia is 2x0)(x"q + 2x0) - x"dx"g la (d-c) = - - / - l ( E coS 00 (278) 2x0Vx"dx"q(x"d + 2x0)(x"q + 2x0) As a check, we may note that for a symmetric-rotor machine, with x"d = x"q = x2, equation 278 reduces to E COS 0n (279) x2 + 2x0 In equation 279 the voltage E cos 00 is twice the applied voltage in the a, 0 circuit, and the reactance (x2 + 2x0) is three times the effective per circuit reactance in the a, 0 circuit as described just ahead of equa-

STATOR OR ARMATURE DECREMENT FACTORS 111 tion 276. Thus ia should be two thirds as big as the component of ib or ic produced by ia and Z'Q. Reference to equations 170 and 236 shows this to be true, as, from equation 170, ib(=! ic) [produced by ia and i0] \ia + i0 but by equation 236 i0 = ia, and so ib(= ic)[produced by ia and i0] = %ia The formula for the proper x2 is now 2x0Vx"dx"q(x"d + 2x0)(x"q + 2x0) x"dx"q(x"d + 2x0)(x"q + 2x0) - x"dx" qd 0g 0 - dg Similarly, from equation 242 the d-c component of ip is Vx"dx"q(x"d + 2x0)(*", + 2x0) - (x"d + 2x0)(x"q + 2x0)] E sm 00 2x0Vx"dx"q(x"d + 2x0)(x"q + 2x0) (281) For a symmetric-rotor machine, with x"d = x"q = x2, equation 281 reduces to E sin 00 ifl (d-c) = (282) 2 The expression within the curved brackets of equation 281 is therefore the reciprocal of the effective x2 for this case. The voltage E sin 00 is I/V 3 times the applied voltage in the /3 circuit, and the reactance x2 is half the effective circuit reactance. Thus ip should be 2/V 3 times as big as the component of ib or ic produced by ip (see equations 170). Summarizing the results obtained so far, we have found that the phase currents ib and ic for a double-line-to-ground fault on phases b and c are given approximately by the equations: E\ ib = -[3x"q cos 6 - \/3(z"9 + 2x0) sin 6]C + %[(x"d + x"q) cos 00 - (x"d - x"q) cos (20 - 60)\A V3~ - [(x"d + x"q + 4x0) sin 00 - (x"d - x"q) sin (20 i (283)

112 DOUBLE-LINE-TO-GROUND SHORT CIRCUIT ET f ic = - -[3x"q cos 0 + \/3V9 + 2x0) sin 0]C + %[(x"d + x"g) cos 00 - (x"d - x"g) cos (20 - 00)]A V3 i r/^" i " where sin 00 - (z"d - x"9) sin (20 - 00)]# J (284) B = t/Ta0W-n) f X d -p c Z d t ^e \ \ z'd + xe a;d + xe / + ^^1 C(i I C -I (i,.,) (285) D = 2[x0(x"d + x",) + z"jx", - a;0(z"d - x",)cos20] For Taa(u.n), see equations 276 and 280 (or 273). For Ta0<-n), see equations 277 and 281 (or 273). For T"d(H-B), see equations 267, 268, and 273. For T'd(n-n), see equations 267, 268, and 273. For xe, see equations 268 and 273. We of course do not have to use the equation for C as given. Instead we can if we wish derive more nearly correct formulas (e.g., see equation 270) for any particular case. Field Current The field current is obtainable from id, the direct-axis component of armature current, exactly as in the case of the line-to-line fault or the three-phase fault, using equations 158, 230, and 231. The current id is denned by equation 217 in terms of ia and ip, and ia and ip are in turn given by equations 241 and 242 with the addition of proper decrement factors A, B, C as defined in equations 285. That is (as we must have realized a little while ago during our derivation of equations 283 and

OPEN-PHASE VOLTAGE 113 284 for ib and ic), A multiplies all the even-harmonic terms of ia, B multiplies all the even-harmonic terms of ip, and C multiplies all the odd-harmonic terms of both ia and ip. FIG. 12. Double-line-to-ground fault. Effect of XQ on open-phase voltage. Comparison with line-to-line fault Open-Phase Voltage The voltage ea across the open phase a is given by where (280) (287)

114 SEQUENTIAL FAULTS or, by equation 173, *o = -Wi + x"q ~ 2x0 + (x"d - x",) cos 26]ia - \(x"d - x"q) sin 20 i& + E cos 6 (288) The maximum open-phase voltage can be either greater or smaller than that for a line-to-line fault, largely depending on the value of x0. Figure 12 shows a comparison of values calculated neglecting all decrements with similar line-to-line values. Sequential Application of Faults The principal effect of the sequential application of faults is on the d-c component of short-circuit current. We shall therefore consider for simplicity only a symmetric-rotor case and shall moreover neglect all rotor resistances. This means that we shall neglect all decrements of the a-c component of current but shall obtain approximately the right answers for the d-c components. With these simplifying assumptions the machine may be represented by the equivalent circuit of Fig. 13, in which 31 Sl^ ic x, * en= -E SIN e eh= -E SIN <e I20') ec= -E SIN (8 +I20') x\ x d x q X0 = 3xn + X! x2 (289) Fia. 13. Simplified equivalent circuit for synchronous machine The propriety of this circuit may be shown directly by applying positive-, negative-, and zerophase-sequence voltages in turn and observing that the currents resulting are correct in all cases. It may also be shown by letting x"d = x"q = Xi in equations 173, whence \l/a = Xiia + E cos 6 ^ = -xiip + E sin 6 (290) ^o = x0io and, by equations 170, "Ao = xiia (x0 xi)i0 + E cos 6 ih = -xiib - (x0 - xi)i0 + E cos (6 - 120) (291) ^c = xiic - (x0 xi)io + E cos (0 + 120) Then, in view of equations 1, the circuit of Fig. 13 results.

SEQUENTIAL APPLICATION OF FAULTS 115 Suppose we consider first the simplest case of x0 xi = 0 and Tl = r0 0. Then the three phase circuits are entirely independent. Consequently a short circuit on phase a from terminal to neutral has no effect on the flux linkages of the circuits of phases 6 or c, and the same is true of short circuits on phases b and c. A short circuit on phase a results in the current, E ia = (cos 6 - cos 0a) (292) Zi where 6a is the value of the machine angle 6 at the instant of short circuit of phase a. Similarly, a short circuit on phase b results in the current, E ib = [cos (6 - 120) - cos (06 - 120)] (293) where 06 is the value of the machine angle at the instant of short circuit of phase b. As an example, we may consider the short circuit of phase a at 6 = 6a = 0, of phase 6 at 6 = 6b = 120, and of phase c at 6 = 0c = 240. Then, it becomes evident that all three phase currents will be completely offset (i.e., will have a maximum d-c component) and that the neutral current in = 3i0 = ia + ib + ic will therefore have a d-c component which is at first in y^) = E/xi, then rises to 2E/xi at a time corresponding to 120 rotation of the rotor, and finally rises to 3E/xi at 6 = 240. The a-c component is at first E/xi and moreover stays at E/Xl until 0 = 240 (because the a-c components of ia and ib are 120 apart in phase so that their sum is equal in magnitude to either), and then decreases to zero. Of course, in an actual case the armature resistance will cause the "d-c" components to decay with time so that the full value of in = 3E/xi will not be realized. It should also be remarked that the principal peculiarities of sequential faults have nothing to do with rotating-machine characteristics as such, all the present discussion applying equally well to static circuits except for the fact that directly at the terminals of a large synchronous generator the circuit resistance is usually considerably lower than for other types of circuits, so that the d-c component is more important. Consider next a short circuit not involving ground, or a short circuit to ground on an ungrounded machine. Then it is evident from Fig. 12 that the zero-phase-sequence reactance does not enter into the problem. The first current appears as a line-to-line fault when terminals 6 and c

116 SEQUENTIAL FAULTS are connected together. From the previous chapter (equation 178) the current is V3E ib = ic = - (sin 6 sin 00) (294) 2xi The maximum d-c component is obtained when 60 = 90 and is equal to V 3/2 or 86.6 per cent of the maximum value of the three-phase fault current. Now suppose that phase a is short-circuited to b and c at a later time, when 6 = 6a. The current in phase a is then E ia = (cos 6 cos 0a) (295) Zl This relation is the same as equation 128 of Chapter 4. It may be simply derived by recognizing that in the simple circuit considered (1) the a-c component of current is independent of the sequence of switching operations and thus is the same as for an ordinary threephase fault, and (2) the d-c component must be just large enough to make the initial current zero. It may also be derived in a more straightforward way by noting that, after the line-to-line fault on phases b and c has occurred, the circuit is in effect a single-phase circuit having a voltage 1.5E sin 6 and an impedance of 1.5xi. That is, the voltage at the short-circuited terminal of phases b and c is by symmetry the average of eb and ee, or l/i[-E sin (6 - 120) - E sin (0 + 120)] = +%E sin 0, whence the voltage from this common terminal to the terminal of phase a is ea %(eb + ec) = 1.5E sin 0. The impedance may be seen from Fig. 13 to be the impedance of phase a in series with the paralleled impedances of phases b and c. When phase a is short-circuited to phases b and c, the current ia produced divides equally between phases 6 and c, so that there are now additional currents, 1E ib = ic = z ia = ~ (cos 0 cos 0a) (296) 2 2xi flowing in phases b and c. The total currents in phases b and c are, by adding the components of equations 294 and 296, E /VZ \/3 1 I ib = I - sin 0 -- sin 00 -- cos 0 + - cos xA 2 2 2 2 ) E/ V3 V3 1 1 \ ic = I sin 0 H sin 00 cos 0 + - cos 0a I xi\ 2 22 2 / (297)

SUMMARY 117 If we short-circuit phase a 90 after the first short circuit, or at 6a = 180, the d-c component in phase a is a maximum, and the total currents in phases 6 and c are Tjt ib = [cos (6 - 120) - 1.366] Xl (298) E ic = [cos (6 + 120) + 0.366] xl The effect is thus to offset the current in phase b more than 100 per cent if the decrement factor is small. There are a great many other possible combinations, including various values of zero-phase-sequence reactance, but it is believed that the two cases discussed are sufficient to illustrate the point and a method of attack. The effect of saliency in combination with sequential short circuiting has been considered in reference 13. Summary In Chapter 6 we have discussed the remaining types of short circuit which may afflict a synchronous machine. Again we have restricted ourselves to the case of small resistance and have thereby been able to avail ourselves of the powerful concept of constant flux linkages, without which the problem would be immeasurably more difficult. We have also found the concepts (considerably broadened, however) of symmetrical components to be indispensable for the attainment of a clear physical picture of the double-line-to-ground fault, particularly in .relation to the rotor decrement factors. In discussing the phenomena of sequential faults we have introduced still a third helpful concept, that of constructing a simple equivalent circuit containing the essentials of the problem and nothing else. The moral we shall draw is that there is no single best way to attack all problems. Instead, we must learn to jump from component system to component system and from concept to concept at the slightest indication of a possible simplification to be gained thereby and follow the path of least resistance to the answer. One interesting point in the double-line-to-ground fault is the necessity for considering two armature time constants and properly combining them.

118 SEQUENTIAL FAULTS Problems 1. Consider a synchronous machine having the characteristics: xd = 1.20 p u x"q = 0.367 xm = 1.20 Xalq = 0.60 x\u = 1.15 iu, = 0.831 Xafd = Xatf = XfU = = 1.00 r = 0.005 x'd = 0.367 tj = 0.0005 x"d = 0.279 m = 0.02 x, = 0.80 n, = 0.04 In order to show the effects of various assumptions regarding negative-sequence reactance X2, calculate and plot the armature and field currents following a double-line-to-ground fault, first by means of the theoretically exact values of initial rotor mmf and final field flux linkages (as was done in Chapter 5 for the line-to-line fault), and then by means of various calculated values of x^. Take x0 = 0.167, 0.279, and 0.367 in turn. 2. Consider a simple representation of a synchronous machine as discussed in the section on sequential application of faults and shown in Figure 13. Calculate the two armature phase currents, t& and ic, for a double-line-to-ground fault through a fault resistance Rf. Plot the magnitudes of the currents, neglecting the d-c components, against Rf/xi for io/xi = 10, 0.5, and 2.0 xi = 0.2 3. Calculate and plot the terminal voltage on the open phase o to ground for the cases of problem 2.

/ SHORT-CIRCUIT TORQUES In this chapter we shall develop some more or less exact, and several approximate, expressions for the torque of a synchronous machine following short circuit. For the cases of balanced short circuit and of synchronizing torques it would be possible to solve for the currents and torques exactly, under our assumption of an ideal synchronous machine. However, just as in the calculation of current, the procedure would be unduly complex, and and we have therefore chosen to start as before with simplified cases neglecting resistance. But now we shall find an important difference. It will not be possible to determine the torques correctly by calculating currents and fluxes neglecting all resistances and then applying decrement factors. This high-handed procedure, which worked so well before, will now be found to overlook entirely the unidirectional components of torque, which arise as i2r losses. Three-Phase Short Circuit with All Resistances Neglected A general expression for torque is given by equation 62 as T = ijd - i^q (62) For a three-phase short circuit from no-load open circuit with all resistances neglected and with constant rotor speed the total currents are given by equations 125 with 3 = 0. e id = - (1 - cos 0 x"d (299) e iq = sin t xq 119

120 SHORT-CIRCUIT TORQUES The flux linkages may be most simply obtained from equations 32 or 112, which with r = 0, u = 1.0, and 5 = 0 become pfa = ed0l = e sin 51 = 0 \O\J\j) ptq + fa = eq0l = e cos 51 = el Solving equations 300 operationally we find for the changes of flux linkage produced by the short circuit: e 1 = -e(l - cos 0 p+1 (301) pe = 1 = -esini p+1 The initial flux linkages are fa0 = +Vq0 = + e (302) 0 = e,,> = 0 so the total armature linkages are, adding equations 301 and 302, fa = e cos t (303) \l/q = e sin t [Note that total current and total flux linkage must be used in the torque formula, and note also that the flux-linkage expressions found in equations 303 do not involve any assumptions about the rotor of the synchronous machine. These equations are valid regardless of the kind of rotor winding or rotor resistance, so that we need know nothing about Xd(p) or xq(p) to find fa and \l/g as long as the armature resistance r is zero.] By equations 299 and 303, the short-circuit torque is e2 e2 T = iqfa idtq = - sin t cos t -\ -- (1 - cos t) sin t (304) X"q X"d or e2 This expression bears a striking resemblance to equations 60 and 84 for the steady-state and transient power angle characteristics if the machine and bus excitations are taken to be equal in those equations.

EFFECT OF ARMATURE RESISTANCE 121 Since all resistances have been neglected, the torque is wholly oscillatory, and there is no overturning tendency. The second harmonic component is small especially in the case of a turbine generator, where the fundamental-frequency component is greatest. This fundamental-frequency component arises from the product of the direct component of id (which decays with the rotor time constants and is proportional to the a-c component of armature current) by the oscillatory \}/q (which, like iq, decays with the armature time constant, as it corresponds with the d-c component of armature current), so it dies away somewhat faster than the rate determined by the armature time constant. However, its initial value may be more than 10 times normal torque for a highspeed turbine generator having an x"d of less than 10 per cent. The practical application of our result appears in the necessity for avoiding the possibility of an amplification of this torque in the shaft of a turbinegenerator set by designing to avoid approaching mechanical resonance at the fundamental frequency. Three-Phase Short Circuit Effect of Armature (Stator) Resistance We have not intended to imply in the previous section that it is legitimate to neglect resistance except in the decrement factors when calculating short-circuit torques, as this is definitely not the case. We shall therefore proceed to take into account first the armature resistance and then the rotor resistances, and then show how we may add into the torque expressions terms taking account of the unidirectional torques which are produced by the resulting i2r losses. With armature resistance included it becomes impossible to solve for the flux linkages independently of x"d and x"q as we did in the previous section. However, we can show that, if we make another assumption, namely: that x"d = x"q, it is unnecessary to determine the change of flux linkages. With negligible rotor resistances the armature flux linkages are given by (306) - Xgtg = The torque is T = iqtd idfrq = ta0iq ~ x"didiq + x"qiqid (307) and, if x"d = x"q, the torque is T = fa0iq = eg0iq = eiq (308)

122 SHORT-CIRCUIT TORQUES Thus it is only necessary to find iq. This is given by the second equation 115 of Chapter 4 [with Xd(p) = xq(p) = x"d = x"q] as (x"dp + r)el ' = *' . (309) or (p + a)1 ^a-^l (310) where r1 Z d .toS or e (p + a)l *, = (312) x"d (p + a j)(p + a + j) By solving this operational expression and multiplying over by e (see equation 308), we find for the torque as a function of time: e2 T = eig = .7T5 5 [~ Vd sin t - r cos i) + r] (313) Z d + 1* This expression for torque shows two things: (1) that for negligible rotor resistance the fundamental-frequency torque decays with the armature time constant (I/a) = Taz, and (2) that the maximum unidirectional or overturning torque occurs when the armature or external load resistance r becomes equal to x"d. The value of r for maximum T may be found by setting dT/dr equal to zero and solving for r. This also shows that e2 1 max (unidirectional) .. ("14) 2z"d which for a subtransient reactance x"d of 10 per cent becomes 5 times normal. Another useful form of the torque expression is e2 [ r ] T = eiq = -"' sin (t - tan"1 a) -f (315) Z L Z\ where Z = Vx"d2 + r2 (316) This form of the torque equation brings out the fact that the overturning torque is of the form i2r, where i = e/Z is the fundamentalfrequency component of armature current which does not decay with the armature time constant and so must decay only as the rotor (average)

EFFECT OF ROTOR RESISTANCE 123 flux decays. We must note that, in calculating the armature i2r loss, it is only the a-c component of current and not the d-c component that contributes to the torque. A final point that should be noted is that the magnitude of the fundamental-frequency component of torque is only very slightly affected by moderate values of armature resistance and thus may be calculated very nearly correctly neglecting resistance. Three-Phase Short Circuit Effect of Rotor Resistance We now consider only the resistances in the rotor circuits. The armature currents are given by equations 122 with e,K> = 0 and w = 1.0 id = 2 1 (317) pe It is evident that, since equations 301 for the changes of flux linkages are still valid, we could also have derived equations 317 directly from the expressions: (318) xq(p) which follow from equations 1 13. We shall assume only a single rotor winding in each axis so that (see equation 150) + xd r (319) + Xq where x'd and x'g are the short-circuit reactances of the direct- and quadrature-axis armature circuits, respectively, and Td and Tg are the

124 SHORT-CIRCUIT TORQUES rotor open-circuit time constants (or the ratio of rotor-coil self-inductance to resistance). The operational impedances 319 may for convenience be put into the forms: (P + a') '- . ay(p) = x' (P + a) (320) where g\yj q . a' Xd xd 1 a= a Td 1 ax'd~ r; 1 P oi Xq T * e X gl n X'a 7", and T'd and T'q may be called the rotor short-circuit time constants. The currents become e (p + a) (321) p(p + /3) Evaluating these operational expressions, we obtain 1 v". " / tt v** j/*. \ i j *d = ^ b + (-a')(1 + O +(*'-;x-2) + (7T7] and ^ . , . ^^^; *9 = ^- e~ft/t H ; r H ; rr (323) or r x'jT2^ i (xd ~ x'd) e~"'j - (xd + x'dT2d) cos t + (xd - x'd)Td sin t ! Xd J (324) a* + x'2T2 x q -\- x q q [- (xg - x'q)Tqe-ft't + (xq + x'qT2q) sin t + (xq - x'q)Tq cos t] (325) To aid in the interpretation of these current equations we may note that the coefficients are simply related to the standstill impedances of

EFFECT OF ROTOR RESISTANCE 125 the machine as measured from the armature terminals at normal frequency. This may be shown as follows. The standstill impedance of the machine viewed from the armature terminals varies from the direct-axis value to the quadrature-axis value, depending on the relative position of the rotor. The direct-axis impedance at normal frequency w = 1, expressed in the usual complex-number notation of a-c circuit theory, is , . . ,. -x'dTd+jxd P3WP) = JXd(3) = , , ._ 1 + jTd x-x'T+x x'dT2d) = Rd+ jXd (326) ,d where (327) xd + x'dT2d and m2 fa - a'd)rd Z2d~*2d + *'d2:T2. (>U>) d xd + x'dT2d '22 x'd2T (330) with similar relations for the quadrature-axis impedance. Substituting these expressions, and the similar quadrature-axis expressions, in the current equations 324 and 325, we find C id = + -5- [RdT'd(ra'1 - Xd cos * + Rd sin t] xd Z2d (331) iq = ~Y l-R^-P'* + Xq sin t + Rg cos t] Zq or, finally, . e eRdT'd id = h ~ e e / Xd\ . H sin [t tan I Z2d Zd \ Rj (332) eRg - ~^'' + cos (t - tan"1 ) Za \ RJ

126 SHORT-CIRCUIT TORQUES The final forms of id and iq are easily recognizable as the currents that would result from application of proper voltages to electric circuits having the impedances Zd(p) = pxd(p) and Zq(p) = pxq(p). In fact, we can see from equations 318 and 301 that id = = pxd(p) 1=H 1=H sin t pxd(p) COSt 1= 1= sin t Zd(p) COSt Zq(p) (333) This concept is of considerable value, since it enables us to compute id and iq by familiar circuit methods without necessarily conforming to DIRECT AXIS Td 5 xffd / 'It QUADRATURE AXIS Zq(p) pXq(p) FIG. 14. Synchronous-machine circuits some conventional way (such as equations 319) of representing the operational impedances. In the present case the final answer (equation 332) is obtainable directly and very simply from equations 333, by noting (1) that the fundamental-frequency component of iq may be written in this form practically by inspection, (2) that the magnitude of the exponential component follows in order to make the initial value of iq zero, (3) that the time constant 1//3' can be seen from the circuit (which simply consists of two magnetically coupled circuits, the primary having no resistance, see Fig. 14), and (4) that id, in addition to having its fundamental-frequency component obtainable by inspection, must be wholly obtainable by integrating iq (and changing subscripts) under the condition that its initial value must also be zero.

DISCUSSION OF THREE-PHASE SHORT-CIRCUIT TORQUES 127 Finally, we must keep in mind that this circuit concept, although not at all dependent on the form of the rotor circuits, depends on the assumption that armature resistance is negligible. After this preliminary skirmish with the armature currents, we proceed to the torque, using again the familiar expression: T = i^d - i^q (62) where the currents are given by equations 331 and the (total) flux link. ages are \l/d = e cos t (303) \l/q e sin t Substituting equations 303 and 331 into equation 62, we find, for the torque: e2/Rd Rq\ /I RdT'd ,\ e2Rq T = -^ + -- + e2 H -- sin * -- e~? cos t /I { \xd 2 \Z2d Z\ \xd Z2d / Z q e2 f 1 / , Rq\ 1 / , Rd\~\ + - sin I 2t + tan"1 -? ) - sin ( 2t + tan"1 ) (334) 2 LZq \ Xq/ Zd \ Xd/J Discussion of Three-Phase Short-Circuit Torques We have now derived three expressions for the torque on a synchronous machine following a three-phase short circuit: Equation 305, in which all resistances are neglected. Equation 315 (or 313), in which all rotor resistances are neglected and in which rotor saliency is also neglected. Equation 334, in which armature resistances are neglected. Let us now examine these expressions in the light of our previous discussions of the physical phenomena occurring in the machine and try to arrive at a practical method of calculating torque for any particular case including both stator and rotor resistances simultaneously. It is evident that the torque may be separated out into four components. 1. A fundamental-frequency component of torque depending principally on the direct-axis subtransient reactance. 2. A unidirectional component of torque proportional to the rotor i2r losses, where i is the a-c component of rotor current or the d-c component of stator current. 3. A unidirectional component of torque proportional to the stator (armature) i2r losses, where i is now the a-c component of armature current.

128 SHORT-CIRCUIT TORQUES 4. A double-frequency component of torque depending on the subtransient saliency. These are discussed in turn below. FUNDAMENTAL-FREQUENCY COMPONENT OF TORQUE This component of torque arises from the product of the direct component of id (or the fundamental-frequency component of armature current) by the fundamental-frequency component of \l/q (or a constant component of armature flux linkage), entirely in equation 305 and principally in equation 334. In equation 334 there is also a small component arising from the product of the direct component of iq and the fundamental-frequency component of \l/d. In equation 315, when armature resistance was considered, we found this same fundamental-frequency torque as the product of the initial direct-axis flux linkage td0 by the fundamental-frequency component of iq. At first sight this may seem disturbing, but we must remember that for convenience we operated on the original torque formula 62 to obtain equation 308 by canceling out certain terms. Thus, all the effect of i^/q was canceled against part of the effect of iqtd. We should not therefore depend on equation 315 for a physical interpretation of the origin of the torque components. The initial magnitude of the fundamental-frequency component of torque is practically proportional to l/x"d. If all resistances are neglected, this relation is seen to be exact (equation 305); if armature resistance is included (equation 315), it becomes proportional to l/Z (where Z = V x"d2 + r2); and, if rotor resistance is included (equation 334), it becomes practically proportional to l/Zd (see equation 328). This last relation is not exact but can be seen approximately as follows: The initial magnitude of the coefficient of e2 sin t in equation 334 is 1 RdT'd Xd Xd / d / d and the initial magnitude of the coefficient of e2 cos t is Rq/Z2q. The combined coefficient of the fundamental-frequency term, which is the square root of the sum of the squares of (Xd/Z2d) and (Rq/Z2q), evidently approaches l/Zd if Rq/Z2q is not too different from Rd/Z2d. Moreover, we are allowed considerable latitude since Rg/Z2g will usually be much smaller than Xd/Z2d. The decrement factor of the fundamental-frequency term is obtainable by noting that its two principal factors are the d-c component of armature flux linkage, which decays according to the armature time

DISCUSSION OF THREE-PHASE SHORT-CIRCUIT TORQUES 129 constant Ta3, and the fundamental-frequency component of armature current, which for usual values of machine resistances decays practically in proportion to the average direct-axis rotor flux linkages. For the synchronous machine with both a field and an amortisseur the decay of the fundamental-frequency component of armature current may be seen in equation 146 of Chapter 4. Thus the fundamental-frequency component of torque may be written approximately as rfund = eh-'/T> \(~ - i-) rt/T" + (4- - -) . \X d X d/ \X d Xd/ or !Tfund = e-'/T8 sin t S e-'/r.i sin t X d Ad 1" -t/T'di + sin t xd. (33G) nt (337) where F is the rotor flux linkage as a fraction of its initial value: X^d Xd / xd \ x'd t \ x'd xd / (338) It should be noted that, if the armature resistance is great enough to affect the impedance, as in equation 315, the fundamental-frequency component of torque will disappear very rapidly. For example, if r/x"d = 0.1, Z/x"d is only 1.005, or still very close to unity, but the time constant of the armature flux linkages is only 10 radians or 1.6 cycles. Thus at a time Y\ cycle after the short circuit (when according to equations 305, 315, or 334 this torque is near its first maximum) it has already decayed to t~T/20 = 0.855 times its initial value of e2/Zd. More nearly exact expressions for torque could be constructed by considering also the quadrature component of rotor average flux as represented by the cos t term of equation 334, but this seems hardly justified in view of the uncertainties involved. FIRST UNIDIRECTIONAL COMPONENT OF TORQUE (ROTOR i2r) This torque is given by the first term of equation 334. This term may be written as Trotor = WdRd + i\Rq) (339) where id = ^ t, = ^- (340) &d %q and is seen to be the average of the losses in the direct- and quadratureaxis rotor circuits. Moreover, it is obtained as the average value of the difference of the products [id (eRd/Z2d) sin t by \/,, = e sin t] and

130 SHORT-CIRCUIT TORQUES [iq = (eRq/Z2q) cos t by ^d = e cos t]. All of these terms correspond to d-c components of armature current and flux linkages or to fundamental-frequency components of rotor currents and fluxes. In effect, it is just as though the armature were supplied with d-c excitation so that the resulting rotor losses must be supplied by mechanical torque, while the stator i2r losses are supplied by the decay of the energy of the stator magnetic field. Since all the factors of the torque decay with the armature time constant Ta3, the torque itself must decay with half this time constant, so that finally 2Z2d if the resistance components are small, this becomes approximately .1 . * If we consider a turbine generator, we may have approximately Zd = Zq = x"d = x"q = 0.10 Rd = Rg = 0.03 Ta3 = 60 radians (about 10 cycles) c = 1.0 Then the rotor i2r torque is Trotor = 3.0 e-'/30 This torque has initially a very large value but rapidly decays. It may in certain cases (e.g., where the shaft natural frequency is rather small) contribute appreciably to shaft torque. As far as the motion of the machine is concerned (e.g., in the calculation of transient stability), this torque is appreciable only for the first 10 cycles, by which time it has decayed to about 0.3, or 10 per cent of its initial value. We may estimate its effect on rotor speed as follows. If we take the rotor moment of inertia M equal to 6000 radians, we may write the torque equation: = -Trotor or 6000pu = -3.0e-'/3 or c O e-"80* J0 -.dHhr= -0.015= -1.5% total change of speed due to rotor i2r losses

DISCUSSION OF THREE-PHASE SHORT-CIRCUIT TORQUES 131 For a waterwheel generator the rotor i2r loss is much smaller. In this case we might have Zd = x"d = 0.25 Zq = x"q = 0.35 Rd = 0.02 Rg = 0.04 Ta3 = 100 radians (about 16 cycles) e = 1.0 whence = 0.323* -'/5 The total change of speed following short circuit, due to this COmponent of torque, may be calculated as before, but now using a smaller rotor inertia of M = 3000, such as may be more nearly correct for a waterwheel generator. In this case, 0 A< = ~ 3000 (50) = ~ or about one third as great as for the turbine generator considered previously. These calculations are of course only approximate, especially in view of the fact that we have in all of our short-circuit calculations assumed constant rotor speed. SECOND UNIDIRECTIONAL COMPONENT OF TORQUE (STATOR /2r) Most of the remarks about the stator i2r losses that follow are based on the analysis made above, neglecting saliency. It can be shown that there is a very small additional unidirectional torque produced by stator resistance in conjunction with saliency (i.e., if x"d 7* x"q) which decays like i2d^. For the present, however, we want to confine ourselves to cases wherein the mathematics is sufficiently simple so as not to' obscure the physical reasoning. As we have remarked previously, it is not possible to see, from the formula T = td0iq, the actual origin of the torque. Instead we must see directly from equation 315 that the unidirectional component arises as i2r, where i is the a-c component of armature current. Indeed, by obtaining only the steady-state solution, it becomes evident that the only torque on the machine arises as a resistance load torque. In general, i2 = i2d + i2q for the direct components of id, iq, or the fundamental-frequency component of armature current, and, if saliency is not neglected, the direct components of id and iq are given by equations 147 to 149 of Chapter 4.

132 SHORT-CIRCUIT TORQUES We may express the d-c component of torque as r.t.tor = i2r = (?d + i\)r (343) where, if the armature resistance r is large (as in the application of a load, rather than a short circuit), the currents are calculated as described in Chapter 4, Short Circuit with Armature Resistance, including equation 157, and, if the armature resistance is small, the component iq becomes negligible (see equations 147-149) and id may be obtained as (see equations 145, 146, and 338). eF . - ^ (344) so that the torque is e2F2 Tstator ^ i r (345) Xd For a large synchronous generator the armature resistance is not likely to be greater than 0.3 per cent so that, for a terminal short circuit at normal voltage, the armature loss torque is not nearly so great as the rotor loss torque. On the other hand, it may persist for a much longer time. If we take for a turbine generator the parameters: r = 0.002 x"d = 0.10 x'd = 0.125 xd = 1.25 e = 1.0 the subtransient value of the stator loss torque is 20 per cent, which decays very rapidly to the transient value of 12.8 per cent and then much more slowly to the steady-state value of 0.128 per cent. For a waterwheel generator with r = 0.003 x"d = 0.25 x'd = 0.333 xd = 1.0 we have Subtransient T = 4.8% Transient T = 2.7% Steady state T = 0.3%

DISCUSSION OF THREE-PHASE SHORT-CIRCUIT TORQUES 133 The fact that the subtransient component of torque disappears very rapidly means that we shall actually obtain the maximum overturning tendency not when r = x"d but when r is made equal to the transient reactance x'd. Since ordinarily the transient reactances of the direct and quadrature axes are widely different, we should include the effect of saliency to find a sufficiently accurate value. Referring to equation 148, the transient torque including the effect of saliency is T' I f-, stator (.< d y' /t i 2\2 (x dxq + r2Y By differentiating T" with respect to r, setting dT'/dr = 0, and solving for r, we find for maximum transient unidirectional torque. The table below shows the values found for r/x'd for various ratios of xq/x'd. 1 1.OO 2 1.06 4 1.30 It therefore appears that x'd rather than x'q is the more important factor in determining the maximum transient load torque. A similar effect may also be noted in the power limit of a salient-pole machine tied to a power system, as in that case also the direct-axis reactance is the more important. DOUBLE-FREQUENCY COMPONENT OF TORQUE The double-frequency component of torque arises from the product of the direct components of armature current and flux linkage, just like the rotor i2r loss torque. It may be see.n from equation 334 to depend in magnitude on the direct- and quadrature-axis standstill (normal-frequency) impedances; its decrement factor is the same as that of the rotor i2r loss torque. If all resistances are small, we may write approximately ) e-2t/Tssin 2t (346) ^ q X d' This component of torque is not particularly significant because it is overshadowed by the much larger double-frequency component which may arise on a line-to-line short circuit, where there is in effect saliency on both stator ancf rotor and all harmonic torques are present. For a

134 SHORT-CIRCUIT TORQUES line-to-line short circuit, the fundamental-frequency torque and the rotor i2r loss torque may be very nearly the same, the second-harmonic torque may be greater, and the stator i2r loss torque may be smaller, than for a three-phase short circuit. Since the armature i2r loss is ordinarily rather small, the line-to-line case may be the more severe. Finally an approximate expression for the complete torque of a salientpole synchronous machine during a three-phase short circuit with all resistances rather small, is .* .* fund l .* rotor izr I stator *'2r ~T~ 1 2d harmonic = (337) + (342) + (345) + (346) p\ -2l/Tai x"d 2\x"d F2r F2r 1/1 1 \ 1 H o + ~ I ) ~ sin 2< _// 2 O\T." 'r"./ X rf _ \X a X d/ J (347) where F is given by equation 338. Expressions valid for either large rotor or stator resistances may be obtained from equations 334 or 313 or, as mentioned at the beginning of the chapter, derived directly from exact solutions for the currents and flux linkages. In nearly all cases, however, it will be sufficient to use the approximate expression above. Line-to-Line Short Circuit In our calculation of the torques resulting from a line-to-line short circuit we shall assume, as before, that the short circuit occurs from no load, that constant (unit) speed is maintained during the period of interest, and that stator and rotor resistances may be neglected in computing the alternating components of torque. Thus resistances will be used only to determine decrement factors and to estimate, as in the three-phase case considered previously, the unidirectional components of torques. GENERAL PROCEDURE21 The synchronous machine is considered to have a main field winding and one amortisseur circuit in its direct axis and one amortisseur circuit in its quadrature axis. Thus both subtransient and transient effects will be obtained. The armature current and rotor mmf are given by the method and equations of Chapter 5, which are reviewed briefly below. The initial

TORQUE 135 values, before the effect of decrements is appreciable, are determined by the open-circuit armature voltage before the fault, the rotor position at the instant of short circuit, and the machine subtransient reactances. The armature mmf is resolved into its direct- and quadrature-axis components and is expressed as odd- and even-harmonic series. The average component of the direct-axis current (or the fundamental component of the actual armature current) induces transient direct currents in the direct-axis rotor circuits; all other harmonic components induce alternating currents in the rotor circuits. Consequently, following the short circuit the d-c component of rotor excitation is increased from its value just before short circuit to a new value which includes the components induced in the rotor windings by the armature current. The decay of the induced d-c components of rotor excitation and that of the odd-series components of armature current are determined by the decrement factors of the direct-axis rotor windings. We assume that the decay of the additional rotor excitation may be represented with sufficient accuracy by two decrement factors, one applying to the decay of the induced direct current in the amortisseur, and the other applying to the decay of the induced component of direct current in the field. A rotor-linkage factor F represents the total average rotor linkages at any time after the short circuit as a fraction of the rotor linkage just before short circuit. The average flux linkage in the short-circuited armature windings is maintained by the induced direct component of armature current. This flux linkage decays with the direct component of armature current. For convenience, an armature flux-linkage factor A is introduced, where A represents the average flux linkage of the short-circuited armature windings as a fraction of the flux linkage of these windings just before the short circuit. The even-harmonic series of the armature current and the odd-harmonic series of the rotor excitation decay in proportion to A. Torque As before, the torque is given by the general expression: T = tdiq - tqid (62) But, by equations 37 and 38, or by equations 24-26, the flux linkages may be expressed in terms of the coil currents as td = Xafdifd + Yq = for a machine with a single amortisseur circuit in each axis. Moreover,

136 SHORT-CIRCUIT TORQUES the changes in rotor excitations may be expressed in terms of the stator currents by equations 215 and 216 as Xaidild = (Xd ~ x"d)id (349) = (xq - x" g)iq if resistance can be neglected. In the quadrature axis all the induced currents are alternating, and there is no initial flux, so that we may write \l/g = (Xq X"g)iq Xqtg = X" gig (350) a result that would have been obvious offhand. We seem to have gone around a circle but shall see that the same procedure will lead to a different result in the case of the direct axis. In the direct axis the initial flux linkage is fa0 so that the total flux linkage is fa = fa0 + (xd - x"d)id - Xdid (351) if rotor resistances can be neglected. However we have seen in Chapter 5 that rotor resistances cannot be neglected in the case of the d-c components of id and \l/d. Equation 351 must therefore be modified so as to give correctly the decrements of the d-c components. The alternating components of fa are always given more or less correctly by equation 351, but the direct components are evidently given approximately by the relations: Initially, subtransient, t"d = fa0 - x"di"d (352) Next, transient, *'d = fa0 - x'di'd (353) Finally, in the steady state, fas = fa0 Xdid, (354) Since we are trying to derive a means of correcting equation 351, we should express these flux linkages in the form: t" , i = fa0 ~ x"di"d (352) t'd = fa0 ~ x"di'd + (x"d - x'd)i'd (355) fa, = fa0 - x"dids + (x"d xd)id, (356) which brings out the fact that, since using the actual current id instead of the average currents i"d, i'd, and i,is will affect only the alternating

TORQUE 137 components of \pd, we may still multiply the armature current id always by x"d if we modify ^m so as to change it into l+do + (x"d - x'd)i'd] bf'do + (x"d - Xd)ids] for the transient condition and into for the steady-state condition In these relations all the currents and flux linkages are only the average (d-c) components. On this basis it may be seen from the discussion in Chapter 5 (from equations 217 to 225) that E id = tdO and Ids x'd + x2 x'd + x2 E ido Xd + x2 xd + x2 (357) (358) Substituting these expressions for average current into the expressions 352, 355, and 356 for average flux linkages we obtain, for the subtransient, transient, and steady-state average flux linkages: +"d = t, dO Vd = "Pdl fas = ^dO - x"di"d (352) ~x"d + x2 ~ x'd + x2 . - x"di'd (359) ~x"d + x2 - Xd + x2 ~ - x"did. (360) The required variation of tda in equation 351, so that the armature current may always be multiplied by x"d to obtain id, is therefore from to to tdO tdO tdO initially (x"d + xa\ \x'd + x2J (X"d + x2\ \xd + x2 / in the transient state in the steady state Therefore the effective value of flux linkage to be added to x"did to obtain \frd may be represented as *J(i-^) L\ x'd + x2 / \x'd e-(/r'*<M> + x2 x"d + x2^ + x2 xd + x2 -i/t' . x d ~r x2 e l" d (M) -|Xd + x2 (361)

138 SHORT-CIRCUIT TORQUES The complete expression for flux linkage fo may therefore be written as - x"did (362) where F is a factor describing the flux-linkage decay, and given by (according to equations 352, 359, and 360) F-.1-" x'd + x2 x2 xd + x2 / xd + x2 (x"d + x2\ l---1 \xd + x2 / (303) The factor F is seen to be of exactly the same form as the F of equation 338 for the rotor flux linkage in case of a three-phase short circuit through an external reactance x2. It will also be recognized as occurring in equation 225 as a factor in the odd-harmonic component of armature current during a line-to-line short circuit, where it represents the decay of the rotor mmf . The expressions 350 and 362 for \l/q and \l/d may now be substituted into the torque equation 62 to obtain T = i^d0F + (x"q - x"d)idiq (364) = i^d0F + (x"g - x"d)id] The currents id and iq may be obtained from ia and ip by equations 217 and 175 (and also adding in the decrement factors as shown in equation 225). 2E(F sin 6 A sin 00) sin 6 id ie sin 6 = x"d + x"q - (x"d - x"q) cos 20 (365) 2E(F sin 6 A sin 00) cos 0 ia = ia cos 0 = - -x"d + x"q - (x"d - x"q) cos 20 where F is given by equation 363 and A = e~'/T<'-o = armature decrement factor Substituting equation 365 in equation 364 and noting that \l/d0 = E, we find, for the torque: 2E2F(F sin 0 - A sin 00) cos 0 T= x"d + x"q - (x"d - x"q) cos 20 2(x"q - x"d)[E(F sin 0 - A sin 00)]2 sin 20 '[x"d + x"q - (x"d - x"q) cos 20]2 (366)

TORQUE 139 For the particular case of zero initial armature flux linkage, 60 0, so that sin 00 = 0, and the torque equation 366 reduces to 2x"qE2F2 sin 26 T" = [x"d + x"q - (x"d - x"q) cos 26}2 Equation 366 may be used to compute the instantaneous torque at any time after short circuit. However we also want to express the torque in terms of its harmonic components, since this may be the best form in which to use it to compute the resulting shaft torques. This may be done by resolving 366 into a Fourier series, whence we obtain 2E2 xd + X [- cos 6 + 36 cos 36 - 5b2 cos 56 + 7b3 cos 76 ---- ] x x"d - x2 . ., A2 - sm2 00 x d + x2 x2 X [sin 26 -2b sin 40 + 362 sin 66 ---- ] (368) where X2 + X"d as in Chapter 5, equations 194 and 203. It is evident from equation 368 that all the odd-harmonic components of torque will disappear as the trapped armature flux linkage decays. On the other hand, only one component of the even-harmonic series will disappear, leaving the even-harmonic series as given by equation 367 for the case of short circuit with zero initial armature flux linkage. The part of the even-harmonic torques proportional to the square of the armature linkages (A2) appears only when the subtransient reactances in the direct and quadrature axes are not equal (see equation 366) . This may be seen also from equation 368, since, if x"d = x"q, then x2 = x"d also. This component is a reluctance torque caused by the variation in equivalent permeance which the rotor presents to the trapped armature flux. Similarly the part of the even-harmonic torque proportional to the square of the rotor flux linkages is caused by the variation in equivalent permeance that the stator presents to the rotor flux. The maximum fundamental-frequency torque, immediately following a short-circuit with maximum armature flux linkage, is 2E2/(x"d + x2).

140 SHORT-CIRCUIT TORQUES This may be compared to the fundamental-frequency torque for a threephase short circuit (from equations 305) of E?/x"d and shows that this component may be just as great for a line-to-line fault as for a threephase fault. In case of a turbine generator, the reactances are low so the torques are large. For example, x"d and x2 may both be only 10 per cent, whence the magnitude of the fundamental component of torque will be 10 times base torque for either type of fault. For this case the three-phase fault has no second-harmonic terms whereas the line-to-line fault (equation 368) has a second-harmonic torque of x2/(x"d + X2) = /4 times the fundamental, or 5 times base torque. For a salient-pole watenvheel generator, with x"d = 0.25, x"q = 0.35, and x2 = 0.30, a comparison of the three-phase and line-to-line torques gives: SHORT-CIRCUIT TORQUES Fund. 2d Harmonic Three-phase 4.0 0.57 Line-to-line 3.64 2.59 It is of interest that the quadrature-axis reactance does not affect the fundamental-frequency component of torque in case of the threephase fault, so that increasing x"q actually increases the maximum torque by adding a second harmonic. As a further illustration of the magnitudes involved, line-to-line shortcircuit torques have been calculated 21 for a salient-pole machine having per-unit constants, as determined by test, of X,l = 0.615, x'd = 0.207, x"d = 0.148 xq = 0.412, x'q = 0.412, x"q = 0.367 'Ta = 40, T'd = 470, T"d = 14.5 Hence, x2 = b= Vx"dx"q = 0.233 and X2 X"d = 0.223 x2 + x"d Figures 15 and 16 show the line-to-line short-circuit torques for this machine with maximum and with zero initial armature flux linkages, respectively, as computed by equation 366. From the standpoint of the design of the machine and of its foundations, it is sufficient to consider the case of maximum initial armature flux linkage, since this includes all the possible torques. For the ma-

TORQUE 141 chine of Figs. 15 and 16 the peak value of torque is nearly 12 times rated torque with maximum initial armature flux linkage, while it is only 4 times rated torque with zero initial linkages. However, it is not necessarily the maximum electrical torque that is the ruling factor, as FIRST THREE CYCLES STEADY STATE 15. Single-phase short-circuit torque, maximum initial armature flux linkages the different harmonic components of electrical torque may be very differently amplified (or reduced), depending on the shaft torsional natural frequencies and on the generator and turbine rotor inertias, before appearing as shaft torque. Note that, although we have considered resistance in computing decrements, we have neglected resistance as far as the magnitudes and phases of the currents and fluxes are concerned. Thus, equations 366, 367, and 368 give more or less correctly only the alternating components of short-circuit torque. The direct or load components of short-circuit torque must be calculated, as in the three-phase case, from the stator and rotor i2r losses. FIRST THREE CYCLES TENTH CYCLE Fig. 16. Single-phase short-circuit torque, zero initial armature flux linkages Just as in the three-phase fault discussed above, only the odd-harmonic components of armature current are to be used in computing the stator i2r loss. On the other hand, both odd- and even-harmonic components now affect the rotor i2r losses. The proper values to use are best seen from the point of view of symmetrical components (i.e., forward and backward rotating mmf's). However, before the discussion of load torque we shall first consider other types of short circuit.

142 SHORT-CIRCUIT TORQUES Other Types of Short Circuit It should be evident that the arguments of the previous section, by which the expressions for direct-axis flux linkage, td = idoF - x"did (362) quadrature-axis flux linkage, *, = -x"qiq (350) and torque, T = iqU/doF + (x"q - x"d)id] (364) were derived do not depend on the kind of short circuit being considered. Equation 364 may therefore be considered as a general equation for the alternating components of torque valid for all faults (including the threephase fault), and we need only list below the formulas for id, iq, F, and A previously derived for each" type of fault in order to obtain the corresponding torque for that type. LINE-TO-LINE ON PHASES b-c (Eqs. 369) 2E(F sin 6 A sin 0O) sin 0 U = H sin 0 = iq = iff cos 0 = x"d + x"q - (x"d - x"q) cos 26 2E(F sin 6 - A sin 0O) cos 6 x"d + x"q - (x"d - x"q) cos 20 ( x"d + x2\ F = I 1 I e ' o -'> \ x'd + x2/ , (x"d + x2 x"d + x2\ _t/T, , X"d + x2 \ X'd + x2 Xd + X2 / Xd + x2 A = e-'/7,.-o x"d + x2 rn mil d(l-l) 1 dO ri mi d(l-lt 1 X'd + X2 X'd + X2 d(l-l) = 1 do Ta{l-l) = Xd + X2 X2 r T = iq [EF + (x"q - x"d)id] (The current ip is taken from equation 175 of Chapter 5, with the appropriate decrement factors.)

OTHER TYPES OF SHORT CIRCUIT 143 LINE-TO-NEUTRAL, ON PHASE a (Eqs. 370) 2E(F cos 6 - A cos 00) cos 0 Id = 1'a COS 0 = x"d + x"q iq = ia sin 0 = f (x"d - x"q) cos 20 -2E(F cos 0 - A cos 00) sin 0 3 + (x"d - x"g) cos 20 x"d + x"q (1 \ x'c d + X2 + Xn e~'" dM X"d + X2 A = e-i/ra(-n, x" 1 di-n - 1 XQ x0 d(i-n) - 1 d0 1 d I-n 1 Xd X'd X0 d (I-n) 1 dO Xd + X2 + XQ 2X2 + XQ 1 o <*-"> ~ "^ - 1 2r + r0 T = iq[EF + (x"q - x"d)id] (The current ia is derived from equations 186 and 185 of Chapter 5.) DOUBLE-LINE-TO-NEUTRAL, ON PHASES b, c E id = ia cos 0 + ip sin 0 = [2x"qF cos 0 (Eqs. 371) - (x"d + x"q)A cos 00 + (x"d - x"q)A cos (20 - 00)] cos 0 E + - [(2x"q + 4x0)F sin 0 - (x"d + x"q + 4x0)B sin 00 + (x"d - x"q)B sin (20 - 00)] sin 0

144 SHORT-CIRCUIT TORQUES iq = ia sin 6 -\- ip cos 6 F = [2x"qF cos 6 - (x"d + x"q)A cos 0O + (x"d - x"q)A cos (26 - 60)] sin 6 + - [(2x"q + ix0)F sin 6 - (x"d + x"q + 4x0)B sin 0O + (x"d - x"q)B sin (26 - 60)] cos 6 where D = 2[x0(x"d + x"q) + x"dx"q - x0(x"d - x"q) cos 26] l/T" '/ d (ll-n) p = /_ x"d + xe\ \ x'd + xe I /x"d + xe x"d + xA \ X'd + Xe Xd + Xe ) Xe = -//7" X d I Xe X<J + Xe X2X0 x2 + x0 x"d + Xe Til _ raw " * d (ll-n) * dO . j X d + Xe X'd + Xe rt ml d (ll-n) ~ * , T, act (ll-n) do x2 + 2x0 r + 2r0 X'2 off (ll-n) r T = iq[EF + (x"q - x"d)id] (The equations for ia and ip used here are obtained from equations 241 and 242 of Chapter 6.)

HARMONIC COMPONENTS OF LINE-TO-NEUTRAL TORQUE 145 THREE-PHASE (Eqs. 372) E id = (F - A cos 0 x "d E io = - A sin t x" Xd Xd A = e-'/rx"d Tff rpif d3 = 1 dO 7X'd T' - T' ^ dZ l M Xd T -X2 ia3 r T = iq[EF + (x"q - x"d)id] It may be noted that, for the three-phase fault, the currents expressed as id and iq are functions of time t (measured from the instant of short circuit) directly rather than of the rotor angle 6 and thus do not depend on the initial rotor angle 6Q (the angle of the direct axis ahead of the axis of phase a at the instant of short circuit). This has been pointed out in Chapter 4 and is a consequence of the complete symmetry of the stator winding. The initial rotor angle determines only the location, and not the magnitude, of the flux trapped in the stator. For all the unbalanced faults, the currents, and consequently the torques depend on the initial rotor angle and have therefore been expressed as functions of 6(6 = 00 + 0. Equations 372 for the three-phase case may be shown to check the alternating components given by equation 347 above. [The equations for id and iq are taken from equations 143 of Chapter 4, with 5 = 0 (no load) and e = E.] Harmonic Components of Line-to-Neutral Torque It has already been pointed out that the principal use of the transient electrical torques that we have been calculating is in the determination of the resulting shaft torques. The mechanical system involved con-

146 SHORT-CIRCUIT TORQUES sists of the generator and at least one, and possibly two, turbines. It is thus an equivalent mass-and-spring system that has one or more natural frequencies of its own. These natural frequencies will ordinarily be so small that components of torque of frequency larger than the second harmonic (120 cycles in a 60-cycle machine) will be appreciably suppressed, and the fundamental-frequency component will be by far the most important term. As the various harmonic components of torque are differently amplified in the mechanical system, it is of interest to have the torques for all types of faults expressed directly in terms of their harmonic components, as has already been done for the threephase and line-to-line cases. The line-to-neutral case is easily written down, since it may be obtained directly from the line-to-line case (equation 368) by simply replacing x"d, x"q, by x"d + x0/2, x"q + *0/2, and replacing 6 by (6 + 90). This was pointed out in Chapter 5 and is also evident from equations 369 and 370 above. The torque for a line-to-neutral fault is thus 2E2 f Xd + X2 + X0 [FA cos 00(sin 6 + 360 sin 36 + 5b20 sin 50 + 7b30 sin 76 H ---- ) / Z2 + *0/2 .8 x"d-X2 \ - [ F2 -- A2 - cos2 00 ) X \ x"d + x2 + x0 x2 + x0/2 / (sin 20 + 260 sin 40 + 3620 sin 60 H ---- )] (373) where x2 - x"d x d ~r x2 + XQ and x2 is defined by equation 204 of Chapter 5. The closed form for the torque, corresponding to equation 366 for the line-to-line case, may also be written down in the same way. A comparison of equations 373 and 368 (or of 370 and 369) shows that the alternating components of torque following a line-to-neutral short circuit are always smaller than those for a line-to-line short circuit at a corresponding initial rotor angle. The line-to-neutral case thus does not have to be calculated so far as these alternating components are concerned. If the machine neutral were grounded through a resistor having a resistance of the same order of magnitude as the machine transient reactance (i.e., 10 or 20 per cent), a large unidirectional component of

UNIDIRECTIONAL COMPONENTS OF TORQUE 147 torque might exist. This kind of resistance grounding is not used ordinarily and will be very exceptional. A similar development of the torque in a Fourier series could be carried out for the double-line-to-neutral fault but has not yet been completed for the general case. The Fourier series for the currents are given in Appendix A. Unidirectional Components of Torque In the determination of the unidirectional components of torque for unbalanced faults, the effect of armature currents other than the fundamental and d-c components will be negjected. This is really on account of the difficulties of calculation but also may be justified by the relatively small magnitudes of these torques as a whole, by the negligible subtransient saliency (which produces the higher harmonics) in the case of turbine generators where the torques are greatest, and by the uncertainty inherent in the available values of the resistances. As stated previously, the even-harmonic currents (now only the d-c components) produce a rotor i2r loss by acting as d-c excitation on the stator just as in the case of the three-phase fault, while the odd-harmonic currents (now only the fundamental) produce not only a stator i2r loss but also a rotor i2r loss. This additional rotor i2r loss is caused by the negative-phase-sequence component of armature current and is best seen from the viewpoint of symmetrical components. Our first step is therefore to derive an expression for unidirectional or average torque in terms of symmetrical components. We have shown in Chapter 2 that the torque T is given in terms of d- and g-axis quantities (i.e., in terms of axes fixed in the rotor) as T = i^d ~ i^g (62) By means of the relations 217 for i and the similar relations for \l/: td = ta cos 6 + \{/B sin 6 (374) tq = ta sin 0 + ^ cos 6 the torque may be expressed in terms of a and /3 quantities (i.e., in terms of axes fixed in the stator) as T = i&a - ** (375) (We note that the expression for torque is exactly the same in terms of stator axes as in terms of rotor axes. Indeed, since the angle 6 does not appear in the result, it is evident that we could have used any other angle just as well in the substitution equations 217 and 374. This brings

148 SHORT-CIRCUIT TORQUES out the fact that the expression for torque is the same in terms of any set of two-phase components, regardless of their position relative to the machine and even of whether or not the co-ordinate axes may have any arbitrary accelerating motion of their own. We may recall the related fact noted in Chapter 2 that the form of the armature voltage equations 32 is also independent of the choice of angle defining the reference axes.) In equation 375 i, t, and the torque itself are all real quantities. Moreover, since for our present purpose we are concerned only with the fundamental components of armature current and with the resulting average torque, both i and ^ are sinusoidally varying quantities. For convenience in calculation and in converting over to symmetrical components (see Chapter 6), equivalent complex exponentially varying quantities will be substituted for the real sinusoidally varying quantities. For example, the complex quantity, i= (o has the real part: i = a cos ut b sin ut and i will be used to represent i in the equations. On this basis the average torque may be written as Tavg = real of [|(i/,*x|a - iaV)] or (376) These forms may be seen to be valid if the dot product is defined, as in Chapter 3, as the sum of the products of the inphase components of the two quantities i and vjf. That is, the average or d-c component of the product of two quantities: (a cos cot b sin ut) times (c cos ut d sin a>0 is %(ac + bd), as may easily be demonstrated, while the dot product of the corresponding complex quantities, (a +#),*"(+#).*" is ac + bd, or just twice the required average value. Similarly, it may easily be shown by trial that the real part of the product of one complex quantity times the conjugate of another, Real of [(a + jb^ut(c - jd)e~ji"] is also ac + bd.

UNIDIRECTIONAL COMPONENTS OF TORQUE 149 The symmetrical components may be denned by equations 256 and 258 of Chapter 6. Substituting the relations: i = ii + i2 .,. ., (258) and $a = ^1 + +2 /ow (377) tye = i(+i - +2) into the average-torque equation 376, we obtain for the torque in terms of symmetrical components: 7,avg = ir0'fi) -!2(/) '378) We have next to derive expressions for the flux linkages t|i, ^2, in terms of the currents ii, 12, and the applied voltages, to see the meaning of this expression 378. For the components of flux linkage due to a short circuit, and for a symmetric-rotor machine (as implied by the neglect of higher harmonics), we may write fd = -x d(p) ii = x(p)id [379) fq = -Xq(p)iq = ~x(p)lq or fd + jfq = -x(p)(i* + jiq) and (380) fd - Jtq = ~x(p)(id ~ jiq) But id = ia cos 0 + 1^ sin 6 iq = ia sin 0 + iff cos 6 or ^ + jV = (ia + jip) cos 6 + (iff - jia) sin 0 = (ia+jiff)*~je Similarly, id -jiq = (ia ~ jip)*+l6 and +d+jiq = (+a + rt/ffh~3'B fd~jfq = (fa -j+ff)l+jB Substituting equations 381 and 382 into 380, fa + jfff = -ejex(p)(ia +jip)*-je or, with p6 = 01, fa + jfff = -X(p - J0))(ia + jiff) (381) (382)

150 SHORT-CIRCUIT TORQUES This relation also holds for the equivalent complex quantities: +a + jtyp = ~x(p - i)(ia + 3k) (383a) By equations 256, the symmetrical-component flux-current relation is therefore \|/! = -x(p - i)ii (383b) Similarly, + - J+/3 = ~x(p + ju) (ia - jip) (384a) and x|/2 x(p + jco)i2 (384b) Thus in terms of symmetrical components and for a symmetric-rotor machine ^i depends only on il, and 4*2 depends only on i2, just as in the case of the d-, g-axis quantities. This is in general not true for the a, /3 quantities, as may be seen from equations 383a and 384a above, except when all rotor resistances are neglected. In this case x(p ju) = x(p + ju) = x" and (again only for a symmetric-rotor machine) v|/a depends only on i, and typ depends only on ip. This may also be seen by reference to equations 173 of Chapter 5. The flux linkages tyi and ^2 given by equations 383b and 384b are only the components due to the fault or load. There are also initial values, which may be computed for the case of initially open circuit as follows: The initial d-, g-axis linkages are: Xafdfifd Vdo = = A = e90 rfd tq0 = 0 = ed0 Therefore the initial a, /3 linkages are "AaO = E cos 6 ^(3o = E sin 0 The equivalent complex quantities ty are tfl) = -JEJe and finally the symmetrical-component linkages are i|/10 = Ej* = E^u+ . n (388) \|/20 = 0 (385) (386) (387)

UNIDIRECTIONAL COMPONENTS OF TORQUE 151 Note here that it is necessary to use the equivalent complex quantities v|/a,3 instead of the real quantities i/x/3 in order to make vfi20 = 0. In the equations for the changes of flux linkages (383 and 384) we did not specify the form of i and \l/, but, since we are concerned only with the fundamental-frequency components and are using equivalent complex quantities, we may let p = ju, whence the ^ i relations become vh = -1(0)1! (389) ^2 = -x(j2u)i2 The total flux linkages are, by equations 388 and 389, (390) v|/2 = -x(j2u)i2 The average torque (equation 378) now becomes T*ve = ii.Wi0 -jXO)ii] + i2.jx(j2u)i2 or, since a;(0) is real, ii .jXO)ii .- 0 and (391) This equation may be interpreted by recalling that for a symmetricrotor machine, r + jx(0) =R!+ jx(G) = Rl + JX, = Zi (392) r + jx(j2u) = Ri + jx(j2u) = R2 + jX2 = Z2 where Zi and Z2 are the conventional symmetrical-component positiveand negative-phase-sequence impedances, respectively, Rl = r, and R2 = r + [real part ofjx(j2u)]. We may also develop initial-voltage relations, just as we developed the initial-flux-linkage relations. Since ed0 = 0 eq0 = E ' eaO = E sin 6 os 6 ei0 = +jEt>9 = j e20 = 0

152 SHORT-CIRCUIT TORQUES Substituting (e10) for and (Z2 - Ri = R2 - Ri + jX2) for [jxti2u)] in equation 391, we obtain ravK = ii.ei0 + i2.(R2 = ii.ei0 + (#2 (394) as a general expression for the average (or unidirectional) component of torque caused by the fundamental-frequency component of armature current. The first term is evidently the total power or i2r loss in the equivalent circuit for whichever type of fault is being considered, while the second term, depending on the rotor resistance at twice-normal frequency and on the negative-phase-sequence component of current, has to be computed separately and added. The torque formula 394 may also be written as ^avg = | U ~Ri + i2 | #2 + io RO + I ia | (#2 #i) = | i, |2/?! + i2 \2(2R2 - fl,) + | i0 \2R0 (395) where RO is the zero-phase-sequence resistance. From the viewpoint of symmetrical components the fundamentalfrequency components of current for the various types of short circuit are:20 LINE TO LINE (396) lQ = 0 LINE TO NEUTRAL ii = i2 = i0 = (397) where ZQ(= RQ + jX0 = 3rg + r + jx0) is the zero-phase-sequence impedance, XQ is the zero-phase-sequence reactance, and rg is the value of any grounding resistor between the machine neutral and ground (see Chapter 5). Note that the fault considered is in general from phase a to ground and not to the machine neutral point.

UNIDIRECTIONAL COMPONENTS OF TORQUE 153 DOUBLE LINE TO NEUTRAL , CIQ fj\ 1~ 7 i_ y Zi2 ~t~ ^0 ia = - , ii = e' (398) Z0 Zi + Z0 H THREE PHASE (399) i2 = i0 = 0 The corresponding unidirectional components of torque are, from equation 395: LINE TO LINE e2i02fl2 rr ;2 or> / \r\f\\ ^avg - | ii | Za - T^^^i^ ^400^ LINE TO NEUTRAL R0) |- _ _._ .a 1^1 T ^2 T ^0 I DOUBLE LINE TO NEUTRAL Z2 + Z0 |2flt + | Z0 \2(2R2 - fit) + Z2 2R0] (401) , _ i ave ZrZ3+ Z2Z0 + 0! THREE PHASE e2R (403) It has become evident from the preceding development that in order to understand and to compute the loss torques due to fundamentalfrequency armature current it is necessary to segregate the current into its positive- and negative-phase-sequence components, rather than into

154 SHORT-CIRCUIT TORQUES its a, 0 or d, q components. However, since we have already computed the a, ft components and since it is, to say the least, convenient to have all terms of the torque formula based on the same quantities, we shall convert the formulas above into a form based on ia, ip. At the same time we shall take into account the fact that all these components of current vary as F (the rotor mmf decay factor), a fact that was not brought out in the symmetrical-component equations 396-399. The fundamental-frequency components of ia, ip, i0 for the various faults have already been obtained in Chapters 5 and 6 as LINE TO LINE (EQ. 218) 2EF sin (00 + 0 X d .+- X" (404) 4 = 0, i0 = 0 LINE TO NEUTRAL (EQ. 226) 2EF ia = - cos (00 + 0 x d + x2+ x0 TO = \ia DOUBLE LINE TO NEUTRAL (EQS. 243 AND 244) 2EF la " x"d + 2x0 + V(*"7 EF (405) X "dX0 X d + XQ + cos (00 +1) 2EF sin (00 + <) (406) " EF - ; - ^- sin (00 + 0 (x2 - x' d)x0 x d H -x2 + 2x0

UNIDIRECTIONAL COMPONENTS OF TORQUE 155 THREE PHASE EF ia = coS (00 + 0 Xd EF i& = sin (60 + 0 (407) x"d i0 = 0 The equivalent complex forms of the currents are: LINE TO LINE ia = 0 (408) . io = 0 LINE TO NEUTRAL X0 DOUBLE LINE TO NEUTRAL EF 0:2 i0 = -ia THREE PHASE EF __ a~ EF .+0 (410) 0=

156 SHORT-CIRCUIT TORQUES From these equations for the fundamental-frequency components of ia, ip, we may derive the symmetrical-component quantities ii, l2, by means of equations 256 of Chapter 6, as LINE TO LINE EF i2 = \ 0- - & x2 EF Xd+ i0 = 0 LINE TO NEUTRAL EF ii = i2 = i0 = | ia = -7 - ; - ^0+0 (413) a; d + x2 + XQ DOUBLE LINE TO NEUTRAL EF xdx2 x"d + x2-i -N X0' EF (414) , X"d + X0-\ -Z2 THREE PHASE EF ,,=, = -, (415) 12 = i0 = 0 Except for the time-varying factors (Fej(9o+'>) these currents may be seen to check the values given in equations 396-399. We may also note that in every case the d-c component of the directaxis current id (the d-c component of iq is zero) is equal to the positivephase-sequence current i leaving out the exponentially varying factor.

D-C COMPONENT OF CURRENT 157 This has already been pointed out for the double-line-to-neutral case in Chapter 6, and may be seen by comparing: For the l-l case, Eq. 219 with Eq. 412 For the Z-n case, Eq. 227 with Eq. 413 For the ll-n case, Eq. 246 with Eq. 414 For the 30 case, Eq. 125 with Eq. 415 The relations derived above, among id; ia, ip; and ii, i2, may be used to evaluate the unidirectional component of torque due to the fundamental-frequency component of armature current by substituting the relations 412-415, instead of 390-399, in equation 395. Unidirectional Component of Torque Due to D-C Component of Current We next investigate the rotor i2r loss component of torque due to the d-c component of armature current (which acts as an exciting current and generates a fundamental-frequency current in the rotor) in combination with the rotor fundamental-frequency resistance. For the three-phase short-circuit case this latter torque component has been given by equation 339 as T = WdRd + i\Rq) (339) where in this equation id ( = E/Zd) and iq ( = E/Zq) are seen from equation 332 to be the magnitudes of the a-c components of id and iq, respectively. These a-c components of id, iq correspond roughly to the d-c components of the actual armature current. Rd and Rq are the effective rotor resistances viewed from the stator, as shown for example by equation 326. From our viewpoint of regarding the d-c armature current merely as an exciting current, it is evident that equation 339 is also valid for the general case, i.e., for any type of fault for which the a-c components of id and iq can be evaluated. In the general case the fundamental-frequency components of id and iq may be derived from the even-harmonic components of ia and ip, which have already been obtained and are listed below for the various types of fault. LINE TO LINE 4=02 ' id = ip sin 6 = ~p ib sin 6 V3 iq ip cos 9 = ib cos 6 2 V3

158 SHORT-CIRCUIT TORQUES By equation 193, 2EA sin 00 [ 1 I id (odd) = sin 6 b sin 6 cos 20 -\ x2 12 5 2EA sin 00 f 1 6 I sin 6 + - sin 6 H 1.2 2 J x2 1.2 and with X2 + X"d 2EA id (fund) = ; sin 00 sin 0 (416) X d i X2 Note that the fundamental-frequency component of id depends not only on the d-c component of ip but also on the second-harmonic component if there is subtransient saliency, i.e., if x2 7* x"d. That is, although we may obtain the d-c component in the rotor (or stator) from the corresponding fundamental-frequency component in the stator (or rotor), the fundamental-frequency component cannot be obtained from the d-c component alone. Similarly, 2EAsin6Qll 1 iq (fund) = COS 0 6 COS 20 COS 0 H x2 \.2 J 2EA sin 00 [~ 1 b = cos 0 cos 0 x2 12 2 2EAx"d sin 00 cos 0 (417) x2(x". -2EA C"q + X2 . sin 00 cos 6 (418) Equations 416 and 418 show that the direct- and quadrature-axis components of current are determined by using the direct- and quadratureaxis reactances, respectively, as positive-phase-sequence reactances just as in the three-phase case but now with the addition of the negativephase-sequence reactance. Thus the a-c component of id is not equal in magnitude to the a-c component of iq even though there is only a single d-c component of armature current.

D-C COMPONENT OF CURRENT 159 LINE TO NEUTRAL By equation 195, 2EA cos 0O ip = 0 . id ia cos 0 = +-ftO COS 0 -ia sin 0 id = x2 + x0 1 but b0 = Therefore, Similarly, 4EA cos 0O 2x2 + x0 x2 %"d x2 + x"d + x0 cos 0 + b0 cos 20 cos 0 + 1 b0 - cos 0 H cos 0 + 22 2^4 ^d (fund) ^ (fund) + + Z"d + X2 + X0 2EA(2x"d + xq) cos 0O cos 0 (419) (2x2 + x0)(x2 + x"d + x0) 2EA cos 0O sin 0 (420) cos 0O sin 0 (421) x"q + x2 + x0 DOUBLE LINE TO NEUTRAL There is no simple yet exact expression for the fundamental-frequency components of id and iq in the general double-line-to-neutral case. An approximate expression is ^d (fund) iq (fund) = + 2EA cos 0O cos 0 2EB sin 0O sin 0 x"d + x2 + 4x0 x"d + x2 2EA cos 0O sin 0 2EB sin 0O cos 0 (422) x"q + x2 + 4x0 + Z2 In Appendix A this relation is compared numerically with a more or less exact expression, and it may be concluded that it should be satisfactory, since when the corresponding torque is important the degree of sub transient saliency is usually small (e.g., as in a turbine generator), whereas when the subtransient saliency is larger (as in a waterwheel generator) the torque is rather small and therefore unimportant.

160 SHORT-CIRCUIT TORQUES THREE PHASE By equations 145, and converting to our present notation, EA id (fund) = COS t X"d (423) EA iq (fund) = + sin t Xq These expressions are approximations to those of equations 332, which include the effect of rotor resistance. Approximate Torque Equations Finally, from all of the expressions for torque and current components discussed previously in this chapter, we may assemble approximate equations for torque for the several types of short circuit. These are derived on the basis that the subtransient saliency is negligible (i.e., that x"q = x2 = x"d and Rd = Rq) and that the short circuit is from no load, open circuit. Note that the alternating components of torque are now given simply by the expression: T = EFiq If saliency cannot be neglected, the more general expressions derived above may be used to obtain each component separately, but this will not ordinarily be necessary. These approximate torque equations are: LINE TO LINE E2 T^ [-FAsm60 cos (00 + 0 + 0.5^2 sin 2(00 + 01 Xd + *?* + *?**f* (424) 2x"d2 x"d2 LINE TO NEUTRAL 2E2 T ^ [FA cos 00 sin (00 + 0 - 0.5/*1 sin 2(00 + 01 2x"d + XQ E2F2(2R2 + R0) (2EA)2Rd cos2 00 + , . rv- + ,n , S (425)

APPROXIMATE TORQUE EQUATIONS 161 DOUBLE LINE TO NEUTRAL T ^ [-FB sin 60 cos (0O + t) + 0.5F2 sin 2(0O + 0] x"d H [FA cos 0O sin (60 + t) - 0.5F2 sin 2(0O + 01 x"d + 2x0 E2F2[(x"d + s0)2r + x20(2R2 - r) + xVgp] -" 2/// x"d2(x"d + 2x0) + IV #A cos 0O V /EB sin 0<>Vl lWw+HHJ& (426) THREE PHASE E2 E2F2r /EA\2 T = FASiat + -7^- + [ )Rd (427) x"d x"d /> 2 '\ n I In these equations, the decrement factors F, A, and B may be obtained from the appropriate equations 369-372. EXAMPLE To illustrate the application of these torque equations, and to show the magnitudes of the torques developed, we consider the case of a turbine generator having constants as follows: x"d = x"q xi = 0.10 per unit xo = 0.10 R2 = 0.03 Rd = 0.04 x'd = 0.14 xd = 1.25 r = 0.002 Bo = 0.003 T"do = 18 radians T'd0 = 2500 radians E = 1.0 (short circuit from no load, normal voltage) These constants are more or less typical. We remark particularly that we have taken xo = x"d. This is done because it is ordinarily recommended that

162 SHORT-CIRCUIT TORQUES sufficient reactance be inserted between the neutral point of a machine and ground to reduce the line-to-neutral short-circuit current to a value no greater than the three-phase short-circuit current. Similarly, R0 is greater than r to allow for resistance in the neutral reactor and in the ground. The short-circuit torques are: LINE TO LINE (00 = 90) T = IOFA sin t - 5F2 sin 2t + 1.5F2 + 4A2 (428) where F = 0.167e-"15 + O.685-"445 + 0.148 A = e-'/5 LINE TO NEUTRAL (00 = 0) T = 6.67FA sin t - 3.33F2 sin 2t + 0.7F2 + 1.784s (429) where F = O.llSe-"16 + 0.675e-"586 + 0.207 A = e-'/43 DOUBLE LINE TO NEUTRAL (00 = 90) T = 10FB sin t - 3.33F2 sin 2t + 0.77F2 + 452 (430) where F = 0.210<r"14 + 0.675e-"366 + 0.115 B = e-"60 THREE PHASE T = IOFA sin t + 0.2F2 + 4A2 (431) where F = 0.286-'/13 + 0.634e-"280 + 0.080 A = e-"50 In each case the initial rotor angle has been chosen so as to maximize the torque. Note that the line-to-line case seems to be the most severe; it not only contains the largest double-frequency and unidirectional components but also has a rotor decrement factor F larger than any except the line-to-neutral case. The large unidirectional component caused by normal-frequency 'current in the rotor has already been remarked during the discussion of the threephase fault at the beginning of this chapter. In the present case it decays with a time constant of 25 radians, or about 4 cycles. In order to show the effect of the neutral reactance on these short-circuit torques, they are also given below for the case ZQ = 0.05, Ro = 0.002, which corresponds to a short circuit to the machine neutral point rather than to ground. The line-to-line and three-phase cases are of course not affected, while the line-to-neutral and double-line-to-neutral cases become:

SUMMARY 163 LINE TO NEUTRAL (00 = 0) T = 8FA sin ( - 4F2 sin 2 + 0.99F2 + 2.56A2 (432) where F = 0.138e-"16 + 0.683-"B18 + 0.179 A = e-' DOUBLE LINE TO NEUTRAL (60 = 90) T = IOFB sin t - 2.5F2 sin 2t + 0.53F2 + 452 (433) where F = 0.231-"14 + 0.665-"338 + 0.104 B = e-'/6 Summary This has been a long and rather involved chapter. It may seem as though we have labored the subject unduly in view of the many approximations made in the final torque equations. However, we have also gained a physical insight into the origin and nature of the torque and have developed the relations necessary to compute it to whatever degree of exactness we wish, even though it has not seemed feasible or desirable to write down complete and explicit general expressions for torque in each case. In particular, the unidirectional components of torque arising from losses in the machine have required considerable discussion and analysis, as contrasted with the alternating components, which were obtained in a completely straightforward manner. Thus, the relatively less important components have required the more careful study. On the other hand, we have again been afforded a demonstration of the importance of using the most suitable components (i.e., the proper point of view) in arriving most quickly and surely at the right answer; the d, g-axis currents for the evaluation of the alternating components of torque and for the rotor .i2r loss corresponding to the d-c armature current, symmetrical components for the evaluation of the i2r loss caused by the a-c component of armature current, and the components ia, ip for the evaluation of the armature decrement factors, particularly in the doubleline-to-neutral case. Perhaps the most striking point brought out in this chapter is that it is at all possible to utilize the currents previously calculated neglecting resistances, to show that these currents give very nearly the correct values of the alternating components of torque even though they miss completely the unidirectional components, and to patch up the result

164 SHORT-CIRCUIT TORQUES by means of the i?r loss concept so as to obtain finally a fairly good answer. This should not make us lose sight of the fact that the torque may be calculated exactly by equation 62 if the currents are calculated exactly, as may be done in the three-phase case for a general salientpole machine, and for any unbalanced-fault case if the machine is assumed to have a symmetric rotor. Problems 1. For the machine of problem 8, Chapter 5, calculate and plot the curves of short-circuit torque, including decrements and iV losses, for line-to-line short circuits at both minimum and maximum initial flux linkages. 2. Neglect saliency in problem 1; that is, use the direct axis parameters for both rotor axes, and recalculate the line-to-line short-circuit torques, using equation 424. Compare the results with those of problem 1. 3. For a salient-pole synchronous machine having, in addition to the field winding, one amortisseur circuit in each rotor axis, develop an expression for torque following the sudden application of a pure resistance load. As in the section on short circuits with high armature resistance in Chapter 4, armature transients (or the p^y and ptq terms in the armature voltage equations) may be neglected. Discuss the result obtained, in relation to the approximate torque formulas of the present chapter. 4. For the machine and system conditions of problem 1, Chapter 4, calculate and plot the torque. As in that problem, neglect all resistances, but calculate the torque both including and neglecting the d-c component of armature current. Note that even though all resistances are neglected there may be a unidirectional component of torque.

8 STARTING TORQUE A synchronous generator is ordinarily started from rest by mechanical shaft torque produced by the prime mover, brought up to approximately synchronous speed, and then synchronized with the power system. However, synchronous motors and condensers must usually be started from rest as induction motors with no field excitation, brought up to as near synchronous speed as possible, and then pulled into synchronism by application of field voltage. This chapter is concerned with the induction-motor torque developed by a salient-pole machine both with and without field excitation. It will be found that, because of the rotor saliency, the torque will pulsate even with no field excitation, so the speed will pulsate also. However, as a first approximation it will be assumed that the speed pulsation may be neglected. The acceleration will also be assumed to be small. Both of these assumptions may be realized by making the rotor inertia sufficiently large. The machine is thus assumed to be running at various constant rotor speeds, at first without field excitation, and the torque is calculated on this basis. We are therefore concerned again, as in Chapter 3, with a steady-state problem, and moreover with a problem involving only balanced-armature-circuit operation. The machine rotor speed is taken as p6 = 1 - s (434) where s is the slip of the rotor with respect to the applied voltage. Note that slip is positive for induction-motor operation and negative for induction-generator operation. In accordance with the convention of Chapter 2, equation 53, if the machine angle (i.e., the angle of the armature open-circuit generated voltage ahead of the bus voltage) is S, the applied voltages in terms of d, q, 0 axis quantities are ed = e sin 5 eq = e cos 5 (53) e0 = 0 165

166 STARTING TORQUE In the present case since the machine is not in synchronism S is continually decreasing at a rate s, and so we may take 5 = -st (435) We do not have to consider the possibility of a phase angle added to 5, because we are concerned only with the steady state and so are at liberty to choose arbitrarily the reference point from which time is measured. ed = e sin st (436) eq = e cos st With no field excitation and at constant rotor speed the armature voltage equations are, from equations 32, ^d = Ptd (1 s)\f/q rid e sin st (437) eq = ptq + (1 i)to nq = e cos st Since we are interested only in the steady-state torque, it will be very convenient to have the applied voltages expressed as exponential functions (whence p may be set equal to js) rather than as trigonometric functions. One way in which this may be done is by defining new quantities, the "forward-rotating" and "backward-rotating" quantities: 1 ^ 1 ^ (438) 1 ib = ^ (id ~ jiq) 1 (to &q) These forward- and backward-rotating quantities bear somewhat the same relations to id and iq as the symmetrical components ii and ig bear to ia and i0 (see equations 256, Chapter 6). They may thus be

STARTING TORQUE 167 regarded as symmetrical components but now referred to axes fixed in the rotor rather than in the stator. They also differ from the usual symmetrical components in the scale factor V 2 and in being, at least for the present, denned in terms of the real sinusoidal id and iq rather than the equivalent complex quantities id and iq. With these definitions equations 437 become ef = IP + j(l - s)]*/ - rif = j ~ S" V2 (439) e _. eb = [p - j(l - s)tyb - rib = -j =(->" V2 The flux-current relations are, from equations 36, Chapter 2, td = Xd(p)id (440) tq = ~Xq(P)iq or, in terms of the new quantities i/y, tb, if, 4, substituting equation 438 in equation 440: "A/ = xs(p)if - xD(p)ib tb = -xD(p)if - xs(p)ib where xs(p) = \[xd(p) + xq(p)} (442) ~ xq(p)] the subscript S standing for "sum," and D for "difference." Now equations 441 are substituted in equations 439 to eliminate the flux linkages and obtain e ~[P + j(l - s)]Mp)v + xD(p)ib] - rif = j eist (443) e _. ~[P - j(l - s)][fe(p)v + xs(p)ib] - rib = -j'"' The voltages j(e/v 2)eJ8' and j(e/ v2)e~.'3d may be applied separately to compute the corresponding components of if and ib. When j(e/JV2)eist is applied, p = +js; when j(e/v2)e~ist is applied, p = -js. Withy(e/V2)eys' applied: e -ixs(js)if - rif - jxD(js)ib = j-~e}st -, 2s)xs(js)ib - rib = 0 (444)

168 STARTING TORQUE whence the currents are V+ = -Kl-2s)xs(js)+jr]~^t (1 2s)xd(js)Xq(js) + j2srxs(Js) + ^ H+ = (1 - 2s)xD(js) ~ ejsi (1 - 2s)xd(js)xq(js) + j2srxs(js) + r2 Similarly, with j(e/\/2)t~3at applied: -j(l - 2s)xs(-js)if - rif - j(l - 2s)xD(-js)ib = 0 e _. +jxD(-js)if +jxs(-js)ib - rib = -j^ze lst and the currents are M _ 9!jA"r~f i\<A ( V2 (1 - 2s)xD(-js) {-~j V(1 - 2s)xd(js)xq(js) - jrs2xs(-js) + r2 Kl-2x)xs(-jx)-jr][-^r4 H- = e (1 - 2s)xd(js)xq(-js) - jrs2xs(-js) + r The total currents are now if = if+ + ifib = h+ + HIt may be observed that if is the conjugate of ib+ ib is the conjugate of if+ so that if = if+ + I&+ ib = ib+ + lf+ = If (445) (446) (447) (448) (449) where a bar superscript means "the conjugate of." It is now further observed that tj is simply the conjugate of if, a result that could have been anticipated from equations 438 and the fact that id and iq are real quantities.

STARTING TORQUE 169 The torque is given by T = *>d - upq (62) To find torque in terms of the / and b quantities, we substitute the relations: 1 id = ;= (if + h) V2 (450) 1 iq = -J 7= (if - h), etc. Vin equation 62 and find T = j(ibfr - irPb) (451) From equations 441, since ib = y then fa = $f, and so the torque becomes T = j(if+f - iffo) (452) This expression gives the whole torque, including the pulsating components of frequency 2s. The flux linkages are given by (from equations 441): fr = xs(js)if+ - xs(-J8)if- xn(js)ib+ - xD(-js)ib(453) tb = -xD(js)if+ - xD(-js)if_ - xs(js)ib+ - xs(-js)ib- = $f If instead of calculatingly and ^b we calculate j^f and j\pb (which may be regarded as flux densities, as they are at right angles to the flux linkages \f), the torque may be written as T = ifjif - if$f = vOVy) + if(Wt) = 2 X real part of [i/OVy)] (454) or T = 2[if(j+f)] (455) where the dot product signifies the sum of the products of the inphase components. The current will be of the form: if = Ifejst + Ib*-j" (456) and the flux density will be of the form: ji/, = BfJst + Bbe-jst (457) where //, Ib, Bf, Bb are merely complex numbers.

170 STARTING TORQUE In terms of these quantities the torque is (by substituting equation 457 and the conjugate of equation 456 in equation 454): T = 2 X real of [///?/ + IbBb + IbBfe^' + IfBb(->2"] (458) The first two terms of equation 458 are constants and give the average torque, whereas the last two give the pulsating components of torque (of twice slip frequency). The average torque is therefore T = 2 X real of (7/B/ + /6B6) or , + Ib.Bb] (459) The quantities //, Ib, Bf< Bb are the coefficients of the exponential factors in the expressions for */+, etc., and are analogous to the ordinary complex-number representation of a-c circuit quantities. They are given by the formulas: -[(1 -28)xa(J8)+j - 2s)xD(js) r2 =__ (1 - 2s)xd(js)xq(js) + j2srxs(js) + r2 (460) fy = -jxft(js)lf - jxD(js)Ib Bb = +jxs(js)Ib + jxD(js)If It now appears that, since the whole torque may be determined from the quantities / and B (by equation 458, or, if only the average torque is required, by equation 459), we do not have to evaluate the conjugate impedances [x(js)] nor do we have to apply the conjugate voltage (e/V2)e~J" in order to determine the torque. Thus, the introduction of the / and b quantities has not led to as much calculation as was indicated by equations 445 and 447. It is also possible to derive a direct expression for the magnitude T2s of the oscillatory component of slip torque given by the last two terms of equation 458. By multiplying out, taking twice the real part, and

RELATION TO APPROXIMATE TORQUE EQUATION 171 defining the angles of the complex quantities / and B by the relations: It = \If\tf" Bf = | Bf (461) we arrive at the expression: 2[|/6|2 \Bfl 2|//!!/*! l i B, Bb 2 + Bb cos (462) Equivalent Circuit Inspection of equations 444 shows us that */+ and ib+ (or // and Ib) may be obtained from an equivalent electric circuit as shown in Fig. 17. In order to derive this circuit we simply divide the second equation 444 by (1 2s), whence it becomes evident that the mutual impedance \SjXD(js). Equations 446 for if- and ib- may be similarly represented, but this is not necessary since we have shown that the torque may be computed without finding these quantities explicitly. The flux densities B/ and Bb as given by the last two equations 460 are also given directly as voltages in the equivalent circuit and are indicated in Fig. 17. Thus if an actual circuit is available the average torque may be measured directly as twice the sum of two wattmeter readings If.Bf and Ib.Bb, according to equation 459. JXq(js) jxq(js) FIG. 17. Equivalent circuit for asynchronous operation of a synchronous machine Relation to Approximate Torque Equation A well-known and common-sense approximate way of calculating the average slip torque of a salient-pole machine is to calculate the torque twice: (1) as if the machine were a symmetric-rotor machine having in both axes the direct-axis characteristics, and (2) as if it were a symmetric-rotor machine having in both axes the quadrature-axis characteristics. The usual induction-motor circuit can be used to calculate each torque, and the d and q axis values are simply averaged. The two

172 STARTING TORQUE circuits for a machine with one amortisseur circuit in each axis are shown in Fig. 18. The torques are given by the rotor i2r/s losses. This approximate method gives fairly good results except possibly in the region of half speed, where it fails to indicate the dip in the torqueJ1TLTL *od *kd J*t DIRECT AXIS QUADRATURE AXIS FIG. 18. Direct- and quadrature-axis equivalent circuits :, + xi = 0.75 ao = 0.61 Xd Xad ~\~ Xl = 1.17 Xq = Xkkd ^ Xad i~ Xkd ^ l,l& Xkkq ^ Xffd = Xad + Xf = 1.297 Xakq = Xafd ^ Xakd ^ Xfkd ^ Xad ^ * .Uo r = 0.0121 or 0.0363 Tkd = 0.0302 r/ = 0.00145 (= 0.0300 with discharge resistor) Tkq = 0.039 speed curve produced by rotor saliency. Its significance may be shown as follows: If we assume that the armature resistance r is negligible and that the speed is appreciably different from one half, then equations 460 become 1 1 _y e + I zd(js)/2\/2 1\e (463)

RELATION TO APPROXIMATE TORQUE EQUATION 173 Thus, the forward current // is determined by the direct- and quadrature-axis circuit impedances in parallel, whereas the backward current Ib does not matter so far.as average torque is concerned because Bb = 0. The average torque with Bb 0 is now T = 2IrBf = X/2//.V2B/ = V2Irje (464) and, if // is written in the form, /(1|1\ V*90's) jxd(js)/ 2\/2 or + T*- (465) it becomes evident that the average torque is simply the average of the rotor-circuit losses in the direct- and quadrature-axis circuits with a voltage equal in magnitude to e applied to each circuit. The two axis impedances are the "operational impedances" as defined in Chapter 2 multiplied over by j, as shown in Fig. 18 and as ordinarily used in the approximate method of calculation or in the usual inductionmotor torque calculations. Thus, the approximate method is shown to be correct if the armature resistance is very small. If the armature resistance is not negligible, this paralleling method can still be used in most cases. A numerical comparison with the exact method is shown below. Also, if there is no saliency, that is, if Xd(js) = xq(js), it is well known that the "approximate" method becomes exact, regardless of the amount of armature resistance. This may be seen from our equations for if and ib (equations 444 and 446) as well as from the equivalent circuit of Fig. 17, since XD(js) = 0. Equations 444 reduce to + r]if+ -j-= *', ib ,V^ where x(js) = xs(js) = xd(js) = xq(js) with similar conjugate expressions for i6_ and if __ As a further check, we may see that equations 460 reduce to If = r Ib = 0 Bf = -jx(js)If

174 STARTING TORQUE and that the torque (equation 458) becomes simply T = 2 X real of /// = 2//.B/ = -2Ir[jx(js)If] = -2 If\2R = - V2If\2R (466) where R is the real part of jx(js), and so is the equivalent rotor resistance at slip s as seen from the stator terminals, as in the usual inductionmotor torque calculation. If the machine is running at exactly one-half synchronous speed, s = 0.5, and the currents become V2 * y I .>'lTl* Yl'c^ I T fooM 'i J21 dw / i^ ^,w"/J (467) /6 = 0 and Bf = j\[xd(js) + xq(js)]If Bb = +j\[xd(js) xq(js)]If The first equation 467 shows that at half speed the current is determined by the direct- and quadrature-axis impedances in series and that the armature resistance may be included in the machine circuits just as is done in the symmetric-rotor induction-motor case. This is also evident from the equivalent circuit of Fig. 17, as the lower mesh now becomes open-circuited. The average torque at half speed is T = 2IrBf = -2\If\2R = - | V2/, 2R (468) where R = real part of %[jxd(js) + jxq(js)] and is the average of the resistances of the direct- and quadrature-axis circuits (not including the armature resistance). Equation 468 shows that the torque may be determined correctly at half speed by putting the d- and g-axis circuits in series with a voltage of magnitude 2e applied and then taking half the sum of the losses in the d- and g-axis rotor circuits. The correctness of this half-speed method is independent of the magnitude of the armature resistance.

COMPARISON OF "EXACT" AND APPROXIMATE METHODS 175 We have thus good reason to suppose that the average slip torque of a salient-pole synchronous machine may be calculated fairly well over most of its speed range by putting the d- and q-axis impedances in parallel (or considering them separately with the same voltage applied to each), whereas at exactly half speed we should put the two impedances in series. Comparison of "Exact" and Approximate Methods In order to show the order of magnitude of the difference between the approximate and the theoretically correct methods, calculations have been made for a typical motor having constants as given on Fig. 18. The basis for drawing the equivalent circuits of Fig. 18 is that we have FIG. 19. Speed-torque curve of a synchronous motor Comparison of exact and approximate methods of calculation: ra = 0.0121. See Fig. 18 for machine constants

176 STARTING TORQUE assumed that the (actually distributed) amortisseur may be represented by an equivalent single winding in each axis, and, in case of the direct axis, that the mutual inductances between field and amortisseur, field and armature, and amortisseur and armature are all of the same magnitude and, in addition, are all equal to xad. The circuits may be simply 0.4 0.6 SLIP FIQ. 20. Speed-torque curve of a synchronous motor Comparison of exact and approximate methods of calculation: ra = 0.0363. See Fig. 18 for machine constants checked by writing down the mesh equations from the diagrams, including the xad (or xaq) branch in each mesh, and comparing them with equations 37-40 of Chapter 2 for p = js. Moreover, it is evident that, if we leave out the armature resistance r and divide through by j, the circuits represent the operational impedances Xd(js) and xq(js). This may be seen also from equations 120 and 121 of Chapter 4. The results of the average slip-torque calculations are shown in Figs. 19 and 20: for a machine with a normal armature resistance of r = 0.0121 in Fig. 19, and for a machine with three times normal armature resistance, r = 0.0363, in Fig. 20. A combination of both armature resistance

AVERAGE TORQUE (d- AND q-AXIS METHOD) 177 and rotor saliency is required to show a difference between the approximate and exact results, but it is of interest to note that the maximum difference is not much greater for the high armature resistance than for the low armature resistance. The degree of saliency is rather small in this synchronous motor, and so the effect is not great, and the approximate method gives fairly good results. In the discussion of reference 3 by Heinz, and also in reference 9, very much greater torque dips are shown. However, in the first case the machine is special in order to emphasize the effect, whereas in the second case it is a generator rather than a motor. Average Torque (d- and q-Axis Method) In deriving the formulas for steady-state current and torque given above, it seemed expedient to change over from the real d- and g-axis quantities to complex /- and 6-axis quantities in order to arrive at these formulas in a completely straightforward manner. However, we have seen that the final equations may be expressed in terms of complex quantities //, Ib, Bf, Bb which bear a strong resemblance to the ordinary complex-number representation of a-c circuit theory. This suggests that it may be possible to calculate the currents and torques fairly simply without ever introducing the /- and 6-axis quantities. Returning to equations 437, ptd (1 s)tq rid = ed = e sin st ptq + (1 s)\l/d riq = eq = e cos st we note that it is possible to calculate id and iq by the usual a-c circuit method of substituting, for the sinusoidal applied voltages, complex exponential voltages the real parts of which are equal to the actual real voltages. Then id and iq will be the real parts of the resulting exponentially varying currents. Then ed and eq become ed = e sin st + je cos st = j(e cos st + je sin st) = jeS*< (469) eq e cos st + je sin st

178 STARTING TORQUE Again following the usual a-c circuit method we may drop the exponential factors and simply use the complex coefficients: ed = je (470) eq = e Since all currents and flux linkages now vary as ej>t, p = js, and we may write (471) = -Xq(js)tq and rid + (1 - s)xq(js)iq = je (472) riq (1 s)xd(js)id = e whence = _ [~r+j(l ~ 2s)xq(js)]je _ (1 - 2s)xd(js)xq(js) +r2 + jrs[xd(js) + xq(js)] (473) - 2s)xd(js)]e *0 = - 2s)xd(js)xq(js) + r2 + jrs[xd(js) + xq(js)] In terms of the real currents and flux linkages, the torque is given by equation 62, but this is no longer the case for the complex currents and flux linkages calculated by equations 473 and 471. We could multiply these complex quantities by e;sj, then take the real part, and finally substitute in equation 62. However, we are now only interested in the average torque. By analogy to the formula for average power in electric circuits the average torque is 7W = Kvik -id .) (474) This equation may be checked by direct substitution as follows: Let the complex quantities be id = a + jb iq = c+ jd fa = e + jf Then equation 474 gives 7W =\(ce + df-ag- bh) (475)

AVERAGE TORQUE (d- AND q-AXIS METHOD) 179 On the other hand, the real currents and flux linkages are id = real of (a + jb)e>" = a cos st b sin st, etc. whence, by equation 62, the instantaneous torque is T = (c cos st d sin st) (e cos st f sin st) (a cos st b sin s<) (g cos st h sin st) = ce cos2 st + df sin2 st (de + cf) sin st cos st ag cos2 st bh sin2 st + (bg + ah) sin s< cos s< = \[(<* + df) + (ce df) cos 2st - (de + cf) sin 2st ag bh (ag bh) cos 2st + (bg -f ah) sin 2s<] (476) the average value of which checks equation 475. We have now incidentally derived a formula (equation 476) for the instantaneous torque directly in terms of the components of the complex-number form of the d- and <?-axis quantities. This does not necessarily imply that it is better to use the d- and g-axis quantities, as the / and b quantities have some advantage in the simplicity of the equations and in the possibility of using an equivalent circuit as discussed in references 9 and 22 and as shown in Fig. 17. To complete the discussion of the direct- and quadrature-axis method we may see here also the effects of half speed and of negligible armature resistance. At half speed, s = 0.5, equations 473 for id and iq reduce to -je (js) + xq(js)] (477) e n.xq(js)] These relations show that the average impedance is effective in limiting both components of currents. By means of equations 471 and 474 and the relation iq jid, it may now further be shown that Tavg = | id \2R where R = real part of %[jxd(js) + jxq(js)] is the average of the resistances of the d- and g-axis circuits, just as in equation 468.

180 STARTING TORQUE For r = 0, equations 473 become e id = r~ (478) Xq(js) while, by equations 471, td e \l/q = je = j\l/d (479) The derivation of the i2R formula for the average torque from equation 474 is obvious. The simplified relations for the condition r = 0 may also be derived directly from equation 437 in the exponential form: jstd - (1 - 8)tq = ed = je ji*l/q + (1 s)\pd = eq = e Without any concern with the values of the operational impedances, equations 480 may be solved to obtain td = e \l/q = je (479) and equations 478 then follow directly from equations 471. Field Excitation We have so far discussed only the torque developed by a synchronous machine when operating at constant, asynchronous speed, with balanced armature voltages applied, and without rotor excitation. This torque is the slip torque or induction-motor torque and is seen to have an average and a twice-slip-frequency component. As the machine approaches synchronism, the field voltage is applied, so there may be a period before synchronism is finally reached during which the machine operates asynchronously but with rotor excitation. Also, following a system disturbance, such as a severe short circuit which is finally cleared, a synchronous machine may find itself operating out of step. It is thus also important to be familiar with the effects of rotor excitation and to have available a method for calculating the complete torque as well as some concept of the nature of the phenomena involved. Since we have assumed constant rotor speed, the system equations are linear and it is possible to calculate the currents and fluxes produced by the rotor excitation acting alone, and then add them to the previously calculated values produced by the stator excitation acting alone to obtain

FIELD EXCITATION 181 the total currents and fluxes for use in the torque equation. On the other hand, we cannot calculate the torques separately (except for the average value of torque) as there are pulsating components of torque produced by the interaction of the two excitations. The steady-state d- and g-axis currents caused by rotor excitation efd acting alone are constant currents so that the differential operator p = 0 in all equations. The required basic relations are just like those for a three-phase short circuit in the steady state. Armature voltages, with e = 0 and p = 0, (481) -(1 - styqo - rid0 = 0 + (1 - s)tM - riqo = 0 Armature flux linkages, with p = 0, ido = G(0)efd - xdido (482) = E Xdido The resulting currents are (1 - sfXqE r + (l - syxdxq r (i - s)rE ? = T, ^td0 (483) (1 s)xq r2 + (1 s)2XdXq In order to add these currents to the previously calculated slip-frequency currents it will be convenient to change them over to the / and b axes, so that 1 (1 - s)E Vo = 7= (*do + Pq0) = 5 2 V 2 r" + (1 syxdxq 1 . (1 - s)E ib0 = 1=. (ido jiq0) = [(! - s)xq + jr] (484) [(1- s)xq - jr] = vo V2^ r* + (1 - s)2x^ The flux linkages may be obtained from equations 482 and 438 as 1 tf'/o = ~^E ~ xs(0)ifo - xD(0)ibo V2 (485) 1 "A60 = ~7^E - xD(Q)yo - XS(0)lbo = "?/0

182 STARTING TORQUE They may be much more simply given by equations 481 and 438 as jr ifo = Vo 1s tbo = + *60 1s (486) The total current if produced by both stator and rotor excitation is, from equations 456 and 484, of the form: if = i/o + Vs* + he-' (487) and the flux density is, from equations 457 and 486 of the form: if/ = B0 + Bf*'s t + Bbe-'" (488) where B0 = jifo = if0 Is The torque is, from equation 454, T = 2 X real of [*/(#/)] = 2 X real of [(i/o + i^~ja t + V8() X(B0 + BfSst + Bbe-*st)] = 2 X real of [lf0Bo + IfBf + IbBb + (ifoBf + ZftBo)^'8' + (lf0Bb + IfB0)t-'s t + IbBft'2" + IfBbe-j2st] (489) Comparing this expression for torque with equation 458, for no rotor excitation, we find that the effects of rotor excitation are to add a term 2if0.B0 to the average torque and to add a component of torque of slip frequency s, but not to affect the torque component of twice slip frequency 2s. Moreover, it is evident that as far as the average torque is concerned the torques produced by rotor and stator excitation may be calculated separately and then added. The torque produced by the rotor excitation alone involves only the steady-state operating characteristics and thus evidently must correspond to the stator i2r loss. Indeed its value is seen to be TT 2if0.B0 = 2if0 Vo = 2 | if0 \2 1s1s. = (ido2 + iq02) -^- (490) 1s

SUMMARY 183 This torque is the speed torque or braking torque and is a load torque tending to slow down the machine. Thus it opposes the inductionmotor torque which is tending to accelerate the machine. FIG. 21. Torques with and without rotor excitation Curve A. Average torque with rotor excitation Curve B. Maximum torque with rotor excitation Curve C. Minimum torque with rotor excitation Curve D. Average torque without rotor excitation Curve E. Maximum torque without rotor excitation Curve F. Minimum torque without rotor excitation Some examples of the torques resulting from simultaneous application of stator and rotor excitation are given in references 23 and 9. Figure 21 is taken from reference 23 and shows especially the rather large pulsating torque with rotor excitation. Summary In this chapter we have discussed various ways of computing the starting torque of a salient-pole synchronous machine. It has been

184 , STARTING TORQUE shown that the theoretically correct torque without rotor excitation has a dip near half speed. This dip is not shown if the torques of the direct and quadrature axes are computed separately (as for a symmetric rotor machine) and averaged. On the other hand, the results of this approximate method are not usually very far off, especially in view of the uncertainty of the calculated rotor-circuit parameters. Motors are designed every day by the approximate method, and sufficiently close agreements with test results are obtained. Rotor excitation has been seen to reduce the effectiveness of the induction-motor torque, and a simple physical picture of this braking torque has been presented. The magnitude of the armature currents has not been discussed in detail, but formulas for the calculation of these currents have been presented incidental to the development of the torque formulas. Problems 1. Calculate and plot by the exact method the average component of starting torque from standstill to synchronism, without field excitation and with unit armature terminal voltage, for a synchronous machine having the characteristics: Xd = 1.5 Xffd = 1.6 xq =1.0 r = 0.01 Xafd =1.3 Ty = 0.01 No amortisseur 2. Repeat problem 1, but use the approximate method; that is, average the torques calculated for two symmetric-rotor machines. 3. Calculate and plot the field current for the machine and conditions of problem 1. 4. Calculate and plot the field current during the starting period for the machine of Fig. 18. Calculate and plot the p u voltage across the field discharge resistor. The no-load, normal-voltage field excitation of the motor is 40 volts; plot the discharge resistor voltage in volts. 5. Calculate and plot the average starting torque for the machine of Fig. 18 but with an external reactance (as of a transformer and system) of xe 0.10 p u on the machine base. The bus voltage is unity. Compare the results with Figs. 19 and 20. 6. Calculate and plot the field current and discharge-resistor voltage for the machine and conditions of problem 5.

VOLTAGE DIP A problem often involved in the application of synchronous generators, particularly in the smaller sizes, is the determination of the largest load that can he applied as a unit to the generator, for example, the largest motor that can be started. The significant factors are the amount by which the generator terminal voltage dips and the time required for the voltage to return to normal, as these factors may affect both the performance of other apparatus which may be connected to the generator terminals and the starting performance of the applied motor load. If the applied load is a simple balanced static impedance it is evident that the analysis required is similar to that for a three-phase short circuit as discussed in Chapter 4. However, if the terminal voltage is eventually to return to normal (or even if it is to stay within reasonable limits) we must include in our calculation the effect of a voltage regulator which will increase the generator excitation by an amount just sufficient to accomplish this purpose. We shall therefore first briefly consider the nature of the regulator and exciter response. A generator voltage regulator is ordinarily a rather complex automatic device, and to study its performance, stability, and response in detail would require another book. In the present case however, we are concerned with rather large changes of terminal voltage, i.e., about 10 or 20 per cent, which we shall assume will cause the voltage regulator to move to its limit (after an appropriately estimated time delay) and stay there. We may then assume that the combination of regulator and exciter (the excitation system) will produce a prescribed change of alternator field voltage as a function of time under the stimulus of this extremely large voltage dip. The armature current and terminal voltage will then be the resultant of the more or less simultaneous effects of application of load and change of field excitation. If the change in terminal voltage is not sufficient to keep the regulator output signal at its limit, a more comprehensive analysis, beyond the scope of this book, is required. 185

186 VOLTAGE DIP As far as the nature of the applied load is concerned, this may in a typical instance be a large induction motor at standstill. The motor will of course eventually start and come up to speed, but its impedance will not change appreciably until it nears its rated speed, which may take a few seconds. In any case, it is evident that, if we assume that the motor does not start, we shall have considered the most severe case from the standpoint of voltage dip and recovery. As discussed in Chapter 4, the only significant effect of the armature transient is to produce a d-c offset in the armature currents, and this d-c offset can never produce any appreciable component of terminal voltage since it is either very small (for small armature resistance) or disappears very rapidly (for large armature resistance) . We shall therefore neglect armature transients. Also the time constants of the amortisseur circuits are ordinarily so small that the voltage is determined principally by the transient (i.e., field) machine characteristics. Subtransient effects may limit the initial voltage drop, but the voltage very rapidly drops further to the value determined by the transient reactance. Moreover, the maximum voltage dip and the recovery time are not appreciably affected by subtransient reactance unless the excitation response is extremely rapid. We shall therefore also neglect the amortisseur. A further assumption that will be made is that the machine speed p6 does not change from its initial value of unity. Actually, there will be a drop of speed unless the power factor of the applied load is zero, but we shall consider this a refinement on the elementary analysis to be presented here. With these assumptions the basic equations are: Armature voltage, from equations 32, ed - -*9 ~ ^ (491) eq = +td riq Armature flux linkage, from equations 36, fa = G(p)efd - xd(p)id The operational impedances are, with no amortisseur, T'dQx'dp + xd -J^r xq(p) = xq, where T'd0 = -

VOLTAGE DIP 187 Finally, G(p) is G(p) = - - (159) + rfd We may recall also that instead of writing G(p)efd as in equations 36 we may use the expression G(p)E, where E = xafdefd/r/d is the armature open-circuit, normal-speed voltage (see equation 48), and now G(p) becomes T, , i d0P + 1 E is also equal to the field voltage measured in per unit of the value required to produce rated armature terminal voltage at open circuit and normal speed. Equations 36 now become td = G(p)E - xd(p)id (493) ^3 = -xqiq We shall consider the applied load as a simple impedance having a resistance R and an inductive reactance X and shall include these values in the operational impedances and the armature resistance. Then we can regard the application of load just like a short circuit as far as the calculation of the armature currents is concerned. That is: xd(p) = xdm(p} + X xg = xqm + X (494) f = rm + R where the subscript m means that the quantities refer to the machine alone. We consider the application of load to a previously unloaded machine and shall calculate first the armature currents caused by the closing of the armature switch and then separately calculate the currents caused by the change of field excitation. Substituting equation 493 in equation 491, with E = 0, we get Xgiq rid = ed0l = 0 (495) xd(p)id - riq = -eq0l From the first equation 495, i, = id (496)

188 VOLTAGE DIP Then the second equation 495 becomes 1 r2] Xd(p) H Ud = eg0i By the first equation 150, r7"d0*> + Zd < r2] I T'MP + i +d*d = or . ld where, as in Chapter 4, r/ rri dz = i dO r2 + z'dZ, ay xdxq and is the effective load field time constant. The solution of equation 499 is id = idi = (i'd ~ ids)e-'/T'd' + id, where and xdxq r iq = id xa (497) (498) (499) (152) (500) (148) (149) (496) We are ordinarily interested only in the magnitude of the terminal voltage. This magnitude is en = i Z (501) where, by equations 33 and 496, Z = \R+jX = (R and = | r + jxq = (r2 + x\ R + j(xqm + X) (502) (503) (504)

VOLTAGE DIP 189 em = id~ zqz x dX + r r and R are both small, this initial terminal voltage becomes X X t'(0 = eo = * dm T .A The initial armature current i'd is given by equation 148. The initial terminal voltage, immediately following the application of the load Z, is therefore (505) (506) X'd X'dm "+" A or '*^ dm X d//i i .i /m \ em = ; eo = eq\) x dmi d (507) X dm ~T X which shows that the initial voltage drop is determined by the machine direct-axis transient reactance for the case of zero power-factor load. On the other hand, if X is very small, as for the application of a resistance load, the initial voltage drop is much smaller; there may even be a voltage rise instead. For example, we may take xqm = 1.0 p u X=0 / / * dm * dm R = 1.0 rm = 0 Then the initial voltage is, by equation 505, 1.414 ew = eqQ (508) Xdm + 1 For any value of transient reactance smaller than 0.414, this expression shows that the terminal voltage of a machine having transient saliency will rise initially on application of a resistance load. Of course, in the steady state (still not considering the action of the voltage regulator), if synchronous reactance is substituted for transient reactance, the voltage will be reduced. As a further example, we may consider a salient-pole machine with a

190 VOLTAGE DIP quadrature-axis reactance of xqm = 0.7, with all other values as in the last example. In this case: 1.22 ej0 = i .n7 / e9 (509) 1 + (J.IX dm and e, > e,0 for x'dm < 0.31. We may expect the range of applied load for which there is an initial rise rather than an initial drop to be greater for a turbine generator than for a salient-pole (waterwheel) generator, because xq is greater and x'd is smaller for the turbine generator. We may note also the related fact that for rated load on a turbine generator the field flux linkage E'q may be smaller than at normal voltage, open circuit. For example, if the rated load is at 0.85 pf then we may take the system parameters as R = 0.85 X = 0.527 Z = 1.0 xqm = 1.2 rm = 0 xq = 1.727 r = 0.85 Zq = 1.925 x'dm = 0.15 x'd = 0.677 whence, by equation 505, e,0 = 1.017e90 or, for ej0 = 1.0, E'q = eq0 = 0.983 Even if the rated power factor is reduced to 0.8, the rated load field flux linkages may still be less than unity if the transient reactance is reduced below 14 per cent. A convenient way'of studying the magnitudes involved in the initial voltage drop is the vector diagram. discussed in Chapter 3 and illustrated in Fig. 7. In principle, we can first assume no change in terminal voltage and calculate the field flux linkage E'q on this basis by the vectordiagram construction. Then we can scale the whole diagram down (or up) until E'q becomes equal to its value just before the application of load.

EFFECT OF VOLTAGE REGULATOR 191 Effect of Voltage Regulator The voltage as calculated above will never return to its initial value and will indeed fall considerably below it in direct proportion to id as given by equation 500, unless the field excitation is increased. We now proceed to find the component of armature current produced by the change in field excitation. Substituting equation 493 in 491 with eA = eq = 0 and with E = AZ?1, we obtain xqiq rid = 0 (510) xd(p)i,i riq = (7(p)A#l or L T'rip + l + 7ql ''' = T'dop + or 1 t (511) T'dzP + 1 X,iXq + TFORM OF AE1 A/?l is the change in field excitation inspired by the drop in armature terminal voltage. We may approximate it in a variety of ways, depending on the magnitudes involved and the desired result. One way, which has been used in reference 24, is to assume that a finite time elapses before any change of excitation occurs and that the excitation voltage then increases linearly until it reaches its ceiling value. If we are interested only in the maximum voltage dip and not in the time to recover to normal, and if the applied load is not too great, we may find (as in reference 24) that we do not have to consider the third phase, wherein the excitation voltage remains stationary at its ceiling value. If, however, the load is very large, the ceiling becomes more important. In that case we may use the third phase or we may (again after a suitable time delay) assume that the excitation voltage rises exponentially as Afl = (Ee - E0)(l - e-t/T.) (512) instead of linearly, where Ec is the ceiling excitation. E0 is the initial excitation. t is now measured from the end of the elapsed regulator operating time, not from the instant of load application. Te is the excitation-system time constant.

192 VOLTAGE DIP This exponential variation of excitation includes the effect of both rate of rise and ceiling voltage in one equation. If the excitation-system time constant is small compared to the machine load field time constant and if the applied load is large (i.e., if Z is small), it may be possible to approximate the change in field voltage by the unit function equation: Atfl = (Ee - E0)l (513) We might now increase the delay time so as to obtain a suitable approximation to the actual average rate of rise of voltage. It is also possible to use an actual response curve and to work out the solution of equation 511 step by step numerically or to fit the actual curve by better analytical approximations. Since the relation between id and iq is the same for field voltage changes as for the armature voltage change, it does not matter whether we first add the two components of id and of iq together and then calculate the voltage by equation 501 or whether we calculate the two components of et \ separately and then add. AE1 = ktl Suppose we first consider the field-voltage equation having a linear rate of rise: AEl = ktl = -1 (514) P Then, by equation 511, Xq r,, < mi I -t'/T'd, i\i /ri R\ 1d = 2d2 = 9 [' "r 1 dz(e "- 1)J U>1t>) where t' = t ti and ti = time delay in the regulating system before the field voltage starts to increase The voltage is, as before, ZgZ (^ \ e'2 = ^d2 V"-tV Xq If we substitute equation 515 into equation 516, it becomes ZaZk [f + T'd,(e-f/r* - 1)] (517) r2

MINIMUM VOLTAGE 193 For t > t\, equation 517 is added to the previously calculated armature voltage en. ZqZ en = Wd ~ ids)*'1 - + ida] or Let and Z0Ze e(i = qAeqQ xdxq + r2 Z2 ^d^ + r XdXq + r2 x'dXq + r2 = #t = tf9 (518) (519) (520) Then, for t > 0, et = e = Kieq0[(K2 - l)rt/r" + 1] and, for t > h (or t' > 0), e( = c(1 + e(2 C(2 = km -h + rdz^-{t-h),Ti' - D] where (521) (522) (523) Minimum Voltage To find the minimum voltage following the sudden switching on of an impedance load, we may set the derivative det/dt equal to zero and solve for the time for minimum voltage. This time may then be substituted into equations 521, 522, and 523 to obtain et (m.m). The minimum voltage will occur after time t = t\, because the terminal voltage en starts downward and has a negative slope at t = t\, while e(2 starts with zero slope and magnitude at t = t\. There is of course no guarantee that the minimum voltage will be reached before the exciter voltage reaches its ceiling. Differentiating equation 522 and solving for the time td at which the terminal voltage is a minimum, we find U = T'dz log* ego(^2 - 1) + ^/Tdt kT'. dz (524)

194 VOLTAGE DIP Substituting this time td into equation 522, we find for the minimum voltage: Kik T'd, log, , + "," - h (525) where Ki and K2 are given by equation 520. This equation has been given in a slightly different notation' as equation 12 of reference 24, where it has been derived under the assumption of negligible load resistance. AE1 = (c- 0)(1 - e-'/T')l We next consider the field voltage equation: Al = (Ec - #0)U - e~'/Te)l (512) 1 = (Ec - E0) 1 (526) Then, by equation 511, xq(Ec - #0) 1 id = id2 = xdxq + r2 (T'dzP + l)(r.P + 1) l (527) or xq(Ec - E0) ids = ""dz where, as before, t' = t - <i By equation 516, ZqZ(Ec - E0) L1+ T'di-Te J T'dz - Te \ xdxq + r L T' ^ or, by equation 520, [TTM ~t'/T'dc rji t'/Te-i 1 (530) rrit ni ^' i dz i e J Thus, for an exponential variation of exciter voltage, et2 from equation 530 is added to en from equation 521, .to obtain the total terminal voltage et. By differentiating the complete expression for terminal voltage, as before, setting det/dt 0, and solving for the time td when the voltage is a minimum, we find fc-J^rl^-tafcl^^-(^T'dz-TeTe e' ^ Ec-E0\ T'dz (531)

REQUIRED EXCITER CEILING 195 This expression 531 for td has been given in reference 25, in the appendix. It may be substituted into equation 522 (i.e., into equation 521 plus equation 530) to obtain the value of minimum voltage. We note here that, since the applied field voltage is now limited to the ceiling value Ec, there may not be a minimum, but the voltage may continually decrease up to t = if the ceiling is not high enough. In fact, we can see from equation 531 that for td < we must have (since eq0 = E0) E, or, approximately, if Te = t\ = r = 0, Ec Xd E0 x'd Xdm + X % dm ~\~ X (533) AE1 = (Ec - Eo)l If we may use the simple expression AE1 = (Ec E0)l, as in those cases where the ceiling is the predominant regulating-system characteristic, we find, by equation 511, xq(Ec - go) n _ r..t./r.^ XdXq + r2 id2 = ^-, ,;-d--r/r*) (534) or, by equation 516, et2= '--,-.,"' (1 -''',") (535) or ZqZ(Ec - E0) n ,-rrr^ XdXq + r2 et2 = K,(Ec - *o)(l - t^'77"") (53G) where if = t t\. Now the minimum voltage obviously occurs either at t = t\ or not at all. Note that this last case may be derived directly from the previous case simply by setting Te = 0. Required Exciter Ceiling If the voltage is to be able to recover to normal without change of load Z, we must have <M > eq0 which, by equations 521 and 530, requires Ec Ki> 1.0 eo

196 VOLTAGE DIP or, by equation 520, the exciter ceiling voltage must be at least Ec xdXn + r2 (537) With r = 0, the ceiling voltage must be at least > (538) eg0 X This is a value greater than that given by equation 533 which merely required that the voltage does not decrease further after its initial transient drop. After we have made sure that we have an adequate exciter ceiling we may solve for the time to full voltage recovery by setting ct = eq or by plotting a complete voltage-time curve. Saturation The effects of magnetic saturation in the machine appear in two aspects. In the first place, we must use a proper value of transient reactance x'dm in calculating the initial and maximum voltage dip. This reactance may be appreciably higher than the value found from a fullvoltage short circuit, because the latter may be reduced by saturation, especially in case of a turbine generator. Since "saturated" and "unsaturated" are only qualitative terms, we define instead a "normalvoltage" and a "normal-current" transient reactance. The normalvoltage transient reactance is for a sudden three-phase short-circuit from an initial full-voltage open-circuit condition; the normal-current transient reactance is for a sudden three-phase short circuit from open circuit and from a voltage that will result in an a-c transient current very nearly equal to rated current. For a turbine generator this voltage may be only 10 per cent, whereas for a salient-pole generator it may be 30 per cent or more. We should use the normal-current reactance for voltage-dip studies. In some cases tests have been made directly by suddenly applying a low-power-factor impedance load to the terminals of an isolated unloaded machine, and an "incremental" transient reactance has been found. In a machine having a very high degree of main-flux-path saturation and practically no leakage-flux-path saturation it is possible for this incremental transient reactance to be even smaller than the normalvoltage short-circuit transient reactance. For the usual machine, however, it will be larger.

EXCITER RESPONSE 197 The second aspect of machine saturation appears in the determination of the final steady-state terminal voltage with ceiling excitation. This may be especially important in the application of large loads, since we must determine how big the exciter has to be, and it will usually be found most economical to use a larger than normal exciter (i.e., one with a higher than normal ceiling voltage). This high ceiling will accentuate the effect of saturation. If the maximum voltage dip has to be limited, we must decrease as much as possible the transient reactance and increase the exciter response; but if the load is extremely large it becomes impracticable to limit the voltage dip to a small value, and we can only try to produce a reasonably fast recovery to normal. We shall find that the exciter should be made larger until saturation severely limits the benefits gained. A procedure that may be used is to prepare one or more volt-ampere characteristics for the alternator for the load power factor in question or for zero power factor, with various ceiling excitations and also with the initial excitation. Then the initial voltage may be calculated from the transient reactance and the various final voltages calculated from the volt-ampere curves. Then an exponential curve may be drawn with no regulator action, connecting the initial and final voltages. Finally, various exponential curves [if we use the simple third regulating system equation, AE1 = (Ec EQ)1] may be drawn connecting a suitable point on the constant-excitation curve with the final voltage for the chosen exciter ceiling. This will enable us to see directly the benefit to be gained and to make a proper choice among various exciter characteristics. We may find cases in which it is necessary for the motor to come up to a reasonably high speed (i.e., for the load impedance to increase) before full voltage can be attained. If only the minimum voltage, and not the voltage recovery, is of interest, this second aspect of saturation is not so important. Exciter Response We have so far merely assumed a time variation of exciter voltage. The actual time variation may be given as a curve of voltage versus time, or the ASA response may be given. The ASA response is defined for the condition of starting from an exciter voltage corresponding to rated load on the alternator and increasing to ceiling as for a maximum regulator signal. The curve may start slowly, then increase in slope, and finally taper off to the ceiling as illustrated in Fig. 22. The ASA response is based on the per unit slope of a straight line drawn from the initial voltage point Ei in such a manner that the volt-time area under it equals the area under the actual voltage response curve for the first

198 VOLTAGE DIP half second. This line is also shown in Fig. 22. Note that it is not the average slope for the first half second. If we read the voltage Ea from our straight line at one second, the ASA response is ASA response = (539) It is of interest to observe that the ASA response gives some weight to the ceiling voltage as well as to the initial rate of rise of voltage. o X IU 0.5 1.0 TIME-SECONDS 1.5 FIG. 22. Exciter voltage rise (showing straight line drawn to determine ASA response) For example, if the voltage rises instantly from Ei to the ceiling value Ec, then the straight line drawn according to the instructions will extend from Ei at t = 0 to Ea = 4(Ec Ei) + Ef at t = 1.0 second, and the ASA response will have the finite value: ASA response = (540) For a ceiling voltage of double the value required for rated-load field, the ASA response will then be 4.0. For a given ASA response different types of regulating systems will behave differently at other initial exciter voltages. If the exciter supplies its own field excitation and the regulator varies its field resistance (self-excited exciter), the rate of voltage rise starting from no-load alternator field voltage may be less than that corresponding to the ASA value; whereas, if a pilot exciter or the regulator itself supplies the main

EXCITER RESPONSE 199 exciter excitation (separately excited exciter), the rate of voltage rise starting from no-load field voltage will be greater than that corresponding to the ASA value. There is also a type of voltage regulator that supplies a voltage in series with a shunt-connected exciter field (buckboost regulator): this type will have the same high rate of response as the separately excited exciter. The voltage-dip estimating charts of reference 24, one of which is included here as Fig. 23, indicate that to 9 10 VOLTAGE DIP PER CENT SYNCHRONOUS REACTANCE -X.. PER CENT 100 120 140 IT ia 19 20 21 22 23 24 25 26 2T 28 TRANSIENT REACTANCE-X'd PER CENT * FIG. 23. Voltage-dip estimating chart Exciter response: k = 2.16 per unit. Approximate ASA response: self-excited 1.0; separately excited 0.6. External reactance: xe = 2.0 per unit obtain a given rate of voltage rise from no load, the separately excited or buck-boost system need have an ASA response of only 60 per cent of that required for a self-excited system. Once we have determined a suitable exciter voltage-time 'curve for the condition in which we are interested, we may convert it over to the per unit system in use for the alternator by dividing the ordinates by the exciter voltage required for rated voltage on the air-gap line at no load. This must be done before we can use the formulas developed in this chapter. The various approximations used above for the exciter volt-time relation may be found simply by fitting them to the actual curve as best we can. For example, the expression AE = kt may be based on a k computed in the same way as is the ASA response, i.e., by drawing a straight line enclosing the same volt-time area as does the actual curve in a

200 VOLTAGE DIP half second. However, if the maximum voltage dip is of primary interest, it may be better to use the estimated time to minimum voltage, rather than a half second, as the time interval considered. We must remember that in our alternator equations time is measured in units of which 2irf units = 1 second (where / is normal alternator frequency in cycles per second), so that either k calculated as above must be divided by 2irf before use in the formulas or the alternator field time constant T'd0 (or T"dz) must be converted to seconds by dividing by 2wf. Voltage Recovery Time Any of the equations for terminal voltage developed in this chapter may be plotted against time, and the time to return to rated voltage read from this curve. However, when the voltage reaches its rated value, the exciter voltage will no longer continue to increase toward its ceiling, and so the equations will no longer be valid. Moreover, in some cases recovery time means the time required for the voltage to enter and to stay within a given voltage band (e.g., 3 per cent). In these cases it is necessary to consider the actual dynamic equations of the regulating system, and this is beyond the scope of the present chapter. Reference 24 shows some of the results of a study of dynamic response made on a differential analyzer, including not only the time lags, limits, and stabilizer of the regulator, but also the saturation characteristics of the alternator and exciter. Effect of Initial Load We have so far considered only the application of inductive load to a previously unloaded and isolated alternator. It is evident that the severity of the voltage dip will be greater for an isolated machine than for a machine connected to a system. However, there may be an already existing local load on an isolated machine which will modify the voltage dip. If the already existing load may be regarded as a fixed inductive impedance, just like the suddenly applied additional load, and if it may also be considered as in the steady state, two things are evident: 1. The initial voltage drop will be smaller than for the no-load case. 2. The final steady-state voltage corresponding to ceiling exciter voltage will be smaller than for the no-load case. One way to calculate the course of the terminal voltage when there is an initial impedance load is as follows: 1. Calculate E'q, the voltage corresponding to the field flux linkage for the initial load.

FIELD CURRENT 201 2. Calculate E0, the initial field excitation voltage (EQ is now no longer equal to eq0). Both (1) and (2) may be calculated from the vector diagram discussed in Chapter 3, by computing Eq from equation 71 and then id, and then finding E'q = Eq (xq x'd)id and E0 = Eq + (xd xq)id. 3. Parallel the initial and suddenly applied load impedances to form the impedance Z = R + jX. 4. Calculate the initial armature current with fixed field flux linkage as xqE'q i'd = x, x 1T2 (541) 5. Calculate the final armature current with fixed field voltage E0 as id, = X" 2 (542) 6. The current is now of the same form as before, namely: id = idi = (i'd - *d.)~l/r'" + ids (500) but the components i'd and ids are now given by equations 541 and 542. The procedure from here on is identical to that for the case of no initial load. We find that e<i may still be expressed as en = KiEMKi - 1)e~l/T<. + 1] (543) where now, however, (xrix* 4- r21 /E'n\ (544) The equations for the component of terminal voltage et2 caused by the change &E in excitation are the same as before: (equations 517, 523, 529, 530, 535, and 536). The procedure outlined above, rather than the direct superposition method of applying e,0l and ed0l to the terminals of the machine, is recommended because it avoids the necessity for paralleling the initial load impedance with the machine impedance, which not only differs in the two axes but also involves the time-derivative operator p, and because it makes the development practically the same as for no initial load. Field Current In addition to the terminal voltage, it may in some cases be of interest to plot the course of the field current. If field current is measured

202 VOLTAGE DIP in terms of the corresponding open-circuit voltage Ei = xaf,flfd, we may write the armature flux linkages as (545) Then, by the armature-voltage equations 491 with zero terminal voltage, Ei xdid riq = 0 (546) +xqiq rid = 0 Solving for Ei, we find 2 (547) *9 and since, by equations 501 and 502, y et = id Z (548) xq then ZqZ It thus appears that the field current Ei is simply a constant times the terminal voltage et, so the shape of the field-current versus time curve will be exactly the same as that of the terminal voltage versus time. In fact, since armature transients and the amortisseur have been neglected, this constant ratio of field current to terminal voltage can be seen at once from the vector diagram. The relation 542, derived for the steady state, is seen to hold instantaneously, and the whole vector diagram merely varies from its initial to its final size without changing in shape. Summary In this chapter we have discussed various elementary aspects of the problem of suddenly applied impedance loads on an isolated synchronous machine. We have seen that the machine transient reactance, the exciter response, and the exciter ceiling voltage may be important factors. When we reach a limiting case wherein the voltage will not recover unless the motor starts or when we want to calculate the complete response curve, we may have to resort to a step-by-step or differential-analyzer solution, but the methods outlined will at least give us an understanding of the phenomena and of the factors involved.

PROBLEMS 203 The necessity for a high-ceiling exciter for the starting of very large loads has been mentioned. We remark that the ratio of ceiling to fullload exciter voltage is not a direct measure of the effective ceiling height; we may need a relatively higher ceiling for a higher-synchronous-reactance (lower short-circuit-ratio) machine. This may be inferred from the fact that the rated-load excitation does not increase in proportion to the excitation required for normal voltage with the low-power-factor starting load, as the machine reactance increases. Another factor which may sometimes enter is the drop in alternator speed as the load is applied. For an exact solution of this problem, we may return to the fundamental equations and solve them step by step or (more likely) by mechanical means. However, since the change of speed is small, it is also possible to attack this problem analytically to obtain an approximate solution. Problems 1. Consider a three-phase, 60-cycle synchronous generator having the characteristics: xd = 1.2 p u x'd = 0.15 Xafd = 1.1 Tf = 0.001 xffd = 1.15 xq = 1.2 p u With the machine initially at normal voltage and open circuit, and with a field voltage of 50 volts, a reactance load of X = 1.0 p u is suddenly applied. (a) Calculate and plot the terminal voltage as a function of time for a linearly rising excitation voltage of the form A = kt. Take k such that the rate of rise is 100 volts per second. (6) Determine the exciter ceiling voltage required to permit the generator terminal voltage to recover to normal. (c) Determine the value of k required to limit the minimum voltage to 85 per cent. 2. In problem 1 the generator voltage drops initially to a value, X em = e9o = 0.87 p u Xd + X and then drops to a somewhat smaller value before finally returning to normal. Consider a machine having the characteristics of that of problem 1 but also having an amortisseur so that it has in addition the characteristics: Xaid xfid 1.1 pu xiig = 1.13 xi.u =1.13 rid = 0.03 Xalq = 1.1 rlq = 0.03

204 VOLTAGE DIP (a) Calculate and plot the terminal voltage for a linearly rising excitation voltage as in part a of problem 1, i.e., with the same value of k. This will require an extension of the methods of this chapter to include the effect of the amortisseur by proper modification of the quantities G(p) and x(p). Do not include the effect of armature transients. (b) Determine the values of k required to limit the minimum voltage t'> 85 per cent, to 87 per cent, to 89 per cent.

APPENDIX A Fourier Series for Currents, and Fundamental-Frequency Components of id and /',, for Double-Line-to-Ground Fault We first record certain mathematical expressions which have been found useful in expanding in Fourier series the expressions for shortcircuit currents, voltages, and flux linkages following unbalanced faults on a synchronous machine. sin 6 A + B - (A - B) cos 20 1 [sin 6 - b sin 30 + V sin 50 - sin 16 -\ ] (A-l) A + VAB' where Vb - Va Vab - A b = r= 7= = -7= (A-2) VB + VA VAB + A cos 6 A + B + (A - B) cos 26 1 A + VAB cos 6 [cos 6 + b cos 36 + b2 cos 56 H ] (A-3) A + B - (A - B)cos 26 1 1 [cos 6 b cos 36 + b2 cos 56 ] (A-4) A + B - (A - B) cos 25 1 [0.5-6 cos 26 + b2 cos 40 ] (A-5) 'AB 205

206 APPENDIX A sin 20 A + B - (A - B) cos 20 2 A + B + 2VAB cos 26 [sin 26 - b sin 40 + b2 sin 60 ] (A-6) 1 A + B - (A - B) cos 26 A- B AB + 2AB X 2(A + B) A+ B + cos 26 - b cos 46 + b2 cos 60 - , (A-7) Referring now to equations 240-242 of Chapter 6, it is evident that for a double-line-to-ground short circuit, A = x"d(x"q + 2x0) B = x"q(x"d + 2xo) (A-8) The current ia of equation 241 can now be expanded in a Fourier series with the aid of equations A-l through A-8 as 2E "x"d + 2x0 + V(x"d + 2x0Xx"q + 2x0)x"d/x"q X [cos 6 - b cos 30 + b2 cos 50 ] E(x"d + x"q) cos 0O + Vx"dx"q(x"d + 2x0)(x"q + 2x0) E(x"d - x"q) cos 0O [0.5 - b cos 20 + b2 cos x"dx"q(x"d + 2x0)(x"q + 2x0) Vx"dx"q(x"d + 2x0)(x"q + 2x0) + x"dx"q + x0(x"d + x"q) x0(x"d - x"q) + where I2[x"dx"q + x0(x"d + x"q)] E(x"d - x"q) sin 0O + cos 20-6 cos 40 + b2 cos 60 x"dx"q + x0(x"d + x"q) + Vx"dx"q(x"d + 2x0)(x"q + 2x0) X [sin 20 - b sin 40 + b2 sin 60 ] (A-9) b= Vx"q(x"d + 2x0) - Vx"d(x"q + 2x0) Vx"q(x"d + 2x0) + Vx"d(x"q + 2x0)

FOURIER SERIES 207 Similarly from equation 242, the )3 component of current ip is 2E x"d + Vx"dx"q(x"d + 2x0)/(*"9 + 2x0) X [sin 0 - b sin 30 + b2 sin 50 ] E(x"d + x"q + 4x0) sin 00 Vx"dx"q(x"d + 2x0)(x"q + 2x0) E(x"d - x"q) sin 60 [0.5 - b cos 26 + b2 cos 40 ] x"dx"q(x"d + 2x0)(z"9 + 2x0) T" ,T" -\- T (T" , 4- 'r'M * at/ q l^ ~Q v* a T^ " qj [Xn(x"d x" ) I h cos 20-6 cos 40 + b2 cos 60 2[x"dz", + x0(z"d + x",)] I E(x"d - x"q) cos 00 x"dx"q + x0(x"d + x"q) + Vx"dx"q(x"d + 2x0)(z", + 2x0) X [sin 20-6 sin 40 + 62 sin 60 ] (A-10) The fundamental components of ia and ip are given in equations 243 and 244 respectively; the d-c components are given in equations 278 and 281. Note that the even-harmonic cosine series are given as the sum of two separate series, one arising from the constant term and the other from the second-harmonic term in the numerators of equations 241-242. In partictular, the d-c components as given by equations 278 and 281 may be obtained by combining these two series. In addition to the fundamental-frequency components (used in the calculation of the rotor decrement factors) and the d-c components (used in the calculation of the stator decrement factors), we are also interested in the fundamental-frequency components of id and iq, as these quantities are used in Chapter 7 in the evaluation of the average component of torque due to the rotor i2r losses. These fundamental-frequency components of id and iq depend on both the d-c and second-harmonic components of ia and ip, as has been shown in the derivation of id and iq in Chapter 7, equations 416-421, for the l-l and l-n short-circuit cases. It is not immediately evident from the rather complicated expressions of equations A-9 and A-10 for ia and ip that any very simple equations for id and iq can be obtained. However, by analogy to the results obtained for the l-l and l-n cases in Chapter 7, and in view of

208 APPENDIX A equations 279 and 282 of Chapter 6, we might expect the currents to be approximately id (fund) = 4 cos 6 + ip sin 6 2EA cos 00 cos 6 2EB sin 00 sin 6 x"d + x2 + 4x0 x"d + x2 ( iq (fund) = ia Sin 6 + 10 COS 6 2EA cos 00 sin 0 2EB sin 00 cos 0 =H x2 where 4 and B are now the decrement factors for the a and 0 components of id and iq, as given in equation 285 of Chapter 6. We can see that this equation is valid for x"d = x"q, and we also can see that it is valid for x0 = (since this case is a l-l fault), so it remains to investigate its error for the general case. We shall take as a numerical example a machine having a rather high ratio of x"q/x"d and two values of xQ as follows: x"d = 0.2 x"q = 0.3 x0 = 0. 1 and x0 = 2.0 For this case and for x0 = 0. 1 the fundamental components of id and iq may be calculated from equations A-9 and A-10 as Eqs. A-9 and A-10 (XQ = 0.1): id (fund) = 2.3854 cos 00 cos 0 0.228.4 sin 00 sin 0 4.544.B sin 00 sin 0 -0.2285 cos 00 cos 0 (A-12) iq (fund) = +2.135A cos 00 sin 0 +0.2284 sin 00 cos 0 -3.712S sin 00 cos 6 -0.2285 cos 00 sin 0

FUNDAMENTAL-FREQUENCY COMPONENTS OF /,, AND i, 209 where the two armature decrement factors A and B have been introduced, as remarked beneath equation A-lI. For the same case, the symmetrical-component equation A-l 1 gives Eq. A-11 fa = 0.1): ^d(fund) = 2.3664 COS 60 COS 6 4.494.B sin 60 sin 6 (A-13) ^(fund) = +2.1164 cos 6n sin 6 -3.6705 sin 00 cos 6 If the decrement factors A and B are assumed to be somewhat alike, we may compare the magnitudes of the currents found by the two methods in the following table: 00 = 90 60 = 0 Eq. id iq id iq A-9andA-10 4.772 3.484 2.613 1.907 A-11 4.494 3.670 2.366 2.116 The approximate values of id are too small; the approximate values of iq are too large. In other words, the approximate formula does not show quite enough effect of the subtransient saliency. On the other hand, the differences are not great in view of the fact that in general the torques corresponding to these currents cannot usually be computed precisely anyway, as the effective rotor resistances may not be known very accurately. If as an extreme case, we use x"d instead of x2 in the equation for id, and x"q instead of x2 in the equation for iq, so that x2 does not appear in the final equations at all, we find 60 = 90 0n = 0 id iq id iq 5.00 3.333 2.50 2.00 Now at 90 things are as we should naturally expect them to be: the approximate value of id is too large and the approximate value of iq is too small, so that the effect of saliency is accentuated. On the other hand, it is rather surprising to note that, for 60 = 0, the approximate value of id is still too small and the approximate value of iq is still too large.

210 APPENDIX A Now we consider the case of x0 = 2.0 and find from the "exact" equations A-9 and A-10 the fundamental components of id and iq as From eqs. A-9 and A-10 (x0 = 2.0): id (fund) = 0.2374 COS 60 COS 6 -0.0244 sin 00 sin 6 -4.5005 sin 00 sin 6 -0.0245 cos 60 cos 0 (A-14) iq (fund) = +0.2344 cos 00 sin 6 +0.0244 sin 00 cos 0 3.6745 sin 00 cos 0 -0.0245 cos 00 sin 0 The approximate equation A-ll gives Eq. A-11 (x0 = 2.0): id (fund) = 0.2374 COS 00 COS 0 -4.4945 sin 00 sin 0 (A-15) iq (fund) = +0.2344 cos 00 sin 0 -3.6705 sin 00 cos 0 Similarly the "extreme" equation similar to equation A-ll but with z2 replaced by x"d (for id), and by x"q (for iq), gives id (fund) = 0.2384 cos 00 cos 0 -5.0005 sin 00 sin 0 (A-16) iq (fund) = +0.2334 cos 00 sin 0 -3.3335 sin 00 cos 0 Summarizing these results as before, we find 00 = 90 So = = 0 Eq. id *9 id t. A-9 and A-10 4.524 3.650 0.261 0.211 A-11 4.494 3.670 0.237 0.234 "Extreme" 5.000 3.333 0.238 0.233

FUNDAMENTAL-FREQUENCY COMPONENTS OF id AND i, 211 Now with x0 = 2.0 the 60 = 90 case is seen to be by far the most important as far as the armature d-c components are concerned. (00 = 90 was also the most important for the previous case of XQ = 0.1.) In this case the approximate equation A-11 gives very good results, whereas the so-called "extreme" equation is not nearly so close. We note further that the approximate equation does not depend on x0 when 00 = 90, and indeed is the same as for a l-l fault. On the whole, we may conclude that the approximate equation A-ll should generally be satisfactory for the purpose of computing the rotor i2r loss torque, as the errors are not great even for the rather extreme cases considered here.

APPENDIX B Torque The component of force in any direction acting between any two groups of current-carrying coils is equal to the rate of change, with respect to displacement in that direction, of the total stored electromagnetic energy of the system.26 Similarly the torque acting on the rotor of an electric machine is equal to the rate of change, with respect to angle, of the total stored magnetic energy in the machine. For a synchronous machine the total stored electromagnetic energy E may be written as (see Chapter 2, equations 5-8) E = \\Xaal1 a + Xbbi2b + Xcei2e + Xffdi2fd + Xlld^ld + Zll^'V H ---- ] + Xabiaib + Xaciaic Xafdiaifd + Xbcibic lfdild + hdilq -\ ---- (B-l) The torque is therefore 2dE 2 r 1 I' . . 3 d . d 1 iaH xab + iaic Xac + 44 Xbc L do do d6 J ,.Id .d d + *o ifd Xafd + iid Xaid + liq Xalq +' L do d6 do [". d d d + ib ifd Xbfd + iid Xbid + iiq Xbiq H ---L do do do J I d d d II + ic ifd xcfd + iid Xdd + iiq xclq -\ ---- (B-2) L op OP d6 j I 212

TORQUE 213 All other terms are zero, since the other inductances are not functions of the rotor angle. The factor % has been introduced (as in equation 50 of Chapter 2) in order to make the per unit torque unity for load corresponding to rated volt-amperes. Now, from the definitions given in Chapter 2, equations 9-16, the rates of change of the inductances with respect to rotor angle are 3 Xaa= 2xaa2 SU1 20 36 3 xbb = 2xaa2 sin 206 = 2xaa2 sin (26 + 120) 36 3 xcc = -2xaa2 sin 26c = -2xaa2 sin (26 - 120) 36 3 xab = +2xaa2 sin (26 + 60) = -2xaa2 sin (26 - 120) 36 3 xbc = +2xa2sin (26 180) = - 2xaa2 sin 20 36 3 xca = +2xaa2 sin (20 - 60) = -2xaa2 sin (20 + 120) 36 3 Xafd = -xafdsm6 36 3 Xbfd = Xa/dsin (0 120) 36 3 Xcfd = -xafd sin (0 + 120) 36 3 Xald = Xaid sin 0 36 3 Xbld = -Xaidsin (0 - 120) 36 3 Xcld = -xaidsin(0+ 120) 36 3 = Xalq COS 0 36

214 APPENDIX B 3 Xbiq = Xaiq cos (6 120) 36 3 Xc'q = -Xalq COS (6 + 120) 66 etc. (B-3) Substituting these relations B-3 into the torque equation B-2, we find T = %xaa2[i2a sin 20 + i2b sin (26 + 120) + i2c sin (26 - 120)] + ixaa2[iaib sin (26 - 120) + ibic sin 26 + icia sin (26 + 120)] s[xafdifd + Xaidiu + ] X [t'a sin 0 + ib sin (e - 120) + ic sin (0 + 120)] - I[XalqHq H ][4 cos e + fb cos (0 - 120) + ic cos (5 + 120)] (B-4) This expression may be considerably simplified by noting that idiq = -f { \[i2a sin 26 + t26 sin (26 + 120) + i2c sin (20 - 120)] + iaib sin (25 - 120) + ibic sin 20 + icia sin (20 + 120)} (B-5) and that id and iq are obviously contained in the last two terms of equation B4. Moreover we may also note that, by equations 25 and 26 of Chapter 2: Xaa2 = \(Xd~ Xq) (B-6) Substituting equations B-5, B-6, and 20 in equation B-4, we obtain T = ~(Xd Xq)idiq + [Xafdtfd + Xal({ild H Vq [Xalqilq + ' " ']*d (B-7) or, by equations 24: T = i,,fc - i,^, (62) which checks the result obtained in Chapter 2 from another point of view. It has been shown in Chapter 7 that the torque may also be obtained from two-phase armature (a, 0) quantities referred to axes fixed in the stator, as T = i&a - ia+f> (375) which may be derived directly from equation 62 by means of the substitutions 217 and 374.

TORQUE 215 It is of interest to note further that a similar expression for torque in terms of armature quantities may also be obtained directly in terms of the armature phase quantities. This may be simply accomplished by substituting equations 169, together with the similar expressions for current, in 375 above, whence the torque is T = - [iattc ~ tb) + ibtta ~ tc) + icttb - *a)] (B-8) o'X/o or T = = [*.(i6 - i) + tb(ic - ia) + tctt. - 6)] (B-9) As noted above, the factor % appears in these expressions in order to make the per-unit torque unity for unit currents and flux linkages.

REFERENCES 1. S. B. CraRYand W. E. Duncan, "Amortisseur Windings for Hydrogenerators," Electrical World, Vol. 115, June 28, 1941, pp. 2204-2206. 2. R. E. Doherty and C. A. Nickle, "Synchronous MachinesI, an Extension of Blondel's Two-Reaction Theory," AIEE Transactions, Vol. 45, 1926, pp. 912926. 3. R. H. Park, "Two-Reaction Theory of Synchronous MachinesGeneralized Method of AnalysisPart I," AIEE Transactions, Vol. 48, 1929, pp. 716-727. 4. M. L. Waring and S. B. Crary, "Operational Impedances of a Synchronous Machine," General Electric Review, November 1932, pp. 578-582. 5. C. Concordia, "Relations among Transformations Used in Electrical Engineering Problems," General Electric Review, July 1938, pp. 323-325. 6. A. W. Rankin, "The Equations of an Idealized Synchronous Machine," General Electric Review, June 1944, pp. 31-36. 7. A. W. Rankin, "Per-Unit Impedances of Synchronous Machines," AIEE Transactions, Vol. 64, 1945, Part I, pp. 569-573; Part II, pp. 839-841.' 8. A. W. Rankin, "The Direct- and Quadrature-Axis Equivalent Circuits of the Synchronous Machine," AIEE Transactions, Vol. 64, 1945, pp. 861-868. 9. A. W. Rankin, "Asynchronous and Single-Phase Operation of Synchronous / Machines," AIEE Transactions, Vol. 65, 1946, pp. 1092-1102. S/ 10. R. E. Doherty, "A Simplified Method of Analyzing Short-Circuit Problems," AIEE Transactions, Vol. 42, 1923, pp. 841-848. 11. C. Concordia, "Rotating Electrical Machine Time Constants at Low Speeds," AIEE Transactions, Vol. 65, 1946, pp. 882-887. 12. AIEE Test Code for Synchronous Machines, No. 503, June 1945. V' 13. F. Von Roeschlaub, "Effect of Sequential Switching on Short-Circuit Currents in Synchronous Machines," General Electric Review, June 1940, pp. 256-261. 14. W. W. Kuyper, "Analysis of Short-Circuit Oscillograms," AIEE Transactions, Vol. 60, 1941, pp. 151-153. 15. C. Concordia and F. J. Maginniss, "Inherent Errors in the Determination of Synchronous-Machine Reactances by Test," AIEE Transactions, Vol. 64, 1945, pp. 288-294. 16. H. C. Stanley, "An Analysis of the Induction Machine," AIEE Transactions, Vol. 57, 1938, pp. 751-757. V 17. R. E. Doherty and C. A. Nickle, "Synchronous MachinesIV, Single-Phase Short Circuits," AIEE Transactions, Vol. 47, 1928, pp. 457-487. 18. Edith Clarke, C. Concordia, and C. N. Weygandt, "Overvoltages Caused by Unbalanced Short Circuits (Effect of Amortisseur Windings)," AIEE Transactions, Vol. 57, 1938, pp. 453-468. ,_ 19. J. B. Smith and C. N. Weygandt, "Double-Line-to-Neutral Short Circuit of an Alternator," AIEE Transactions, Vol. 56, 1937, pp. 1149-1155. 216

BIBLIOGRAPHY 217 20. Edith Clakke, Circuit Analysis of A-C Power Systems, Vol. ISymmetrical and Related Components, John Wiley & Sons, 1943. 21. C. A. Nickle, C. A. Pierce, and M. L. Henderson, "Single-Phase ShortCircuit Torque of a Synchronous Machine," AIEE Transactions, Vol. 51, 1932, pp. 966-971. 22. Gabriel Kron, "Equivalent Circuit of the Salient-Pole Synchronous Machine," General Electric Review, Vol. 44, 1941, pp. 679-683. 23. A. H. Lauder, "Salient Pole Motors out of Synchronism," AIEE Transactions, Vol. 55, 1936, pp. 636-649. 24. H. C. Anderson, Jr., "Voltage Variation of Suddenly Loaded Generators," General Electric Review, August 1945, pp. 25-33. 25. W. C. Johnson, Jr., "Starting Induction-Type Motors from Systems of Limited Capacity," General Electric Review, December 1930, pp. 712-722. 26. J. C. Maxwell, Electricity and Magnetism, Vol. II, Chapters VI and VII, Oxford University Press, 1892. Bibliography This bibliography is by no means complete, as it contains merely some of the articles that the author has found to be of interest. On the other hand, it is not confined to the scope of the present volume. Instead, it contains treatments of many other topics, including background material, determination of machine parameters, additional applications of the theory developed here, and some of the more recent extensions of the theory. Papers already listed among the references are not included. 1. A. Blondel, Synchronous Motors and Converters, Part 3, Chapter I, McGrawHill Book Company, 1913. 2. Ft. E. Doherty and O. E. Shirley, "Reactance of Synchronous Machines and Its Applications," AIEE Transactions, Vol. 37, Part 2, 1918, pp. 1209-1297. 3. A. Blondel, "The Two-Reaction Method for Study of Oscillatory Phenomena in Coupled Alternators," Revue ginirale de I'electricity, Vol. 13, February, March, 1923, pp. 235-251, 515-531. 4. F. Kade, "Influence of the Damping Torque Winding on Short-Circuit Synchronous Machines," Archiv fur Elektrotechnik, Vol. 12, June 1923, pp. 345-359. 5. R. E. Doherty and C. A. Nickle, "Synchronous Machines IISteady-State Power-Angle Characteristics," AIEE Transactions, Vol. 45, 1926, pp. 927-942. 6. R. E. Doherty and C. A. Nickle, "Synchronous Machines IIITorque-Angle Characteristics under Transient Conditions," AIEE Transactions, Vol. 46, 1927, pp. 1-18. 7. A. R. Stevenson and R. H. Park, "Graphical Determination of Magnetic Fields," AIEE Transactions, Vol. 46, 1927, pp. 112-136. 8. R. W. Wieseman, "Graphical Determination of Magnetic Fields; Practical Application to Salient-Pole Synchronous-Machine Design," AIEE Transactions, Vol. 46, 1927, pp. 141-154. 9. P. L. Alger, "The Calculation of the Armature Reactance of Synchronous Machines," AIEE Transactions, Vol. 47, 1928, pp. 493-512. 10. A. Mandl, "Single-Phase Short Circuits of Three-Phase Generators," Archiv fur Elektrotechnik, Vol. 19, March 1928, pp. 485-513. 11. R. H. Park and B. L. Robertson, "The Reactances of Synchronous Machines," AIEE Transactions, Vol. 47, April 1928, pp. 514-535.

218 REFERENCES 12. R. H. Park, "Definition of an Ideal Synchronous Machine and Formula for the Armature Flux Linkages," General Electric Review, Vol. 31, June 1928, pp. 332334. 13. T. M. Linville, "Starting Performance of Salient-Pole Synchronous Motors," AIEE Journal, Vol. 49, February 1930, pp. 145-147. 14. Gabriel Kron, "Generalized Theory of Electrical Machinery," AIEE Transactions, Vol. 49, April 1930, pp. 666-685. 15. R. E. Dohertt and C. A. Nickle, "Synchronous Machines VThree-Phase Short Circuits," AIEE Transactions, Vol. 49, 1930, pp. 700-714. 16. Y. Ikeda and M. Mori, "Single-Phase Short Circuit of Synchronous Machine," Zeitschrift fur angewandte Mathematik und Mechanik, Vol. 11, August 1931, pp. 274-284. 17. S. H. Wright, "Determination of Synchronous-Machine Constants by Test," AIEE Transactions, Vol. 50, December 1931, pp. 1331-1350; AIEE Journal, Vol. 50, December 1931, p. 961. 18. B. L. Robertson, "Synchronous-Machine Reactance Measurements," General Electric Review, Vol. 35, February 1932, pp. 126-130. 19. F. A. Hamilton, Jr., "Field Tests to Determine Damping Characteristics of Synchronous Generators," AIEE Transactions, Vol. 51, September 1932, pp. 775779; General Electric Review, Vol. 35, July 1932, pp. 384-388. 20. S. B. Crary and M. L. Waring, "Torque-Angle Characteristics of Synchronous Machines following System Disturbances," AIEE Transactions, Vol. 51, September 1932, pp. 764-773. 21. L. P. Shildneck, "Synchronous-Machine Reactances, a Fundamental and Physical Viewpoint," General Electric Review, Vol. 35, November 1932, pp. 560565. 22. R. H. Park, "Two-Reaction Theory of Synchronous MachinesII," AIEE Transactions, Vol. 52, 1933, pp. 352-355. 23. W. M. Hanna, "Uses of Synchronous-Machine Quantities in System Studies," General Electric Review, Vol. 36, March 1933, pp. 116-128. 24. S. B. Crary, L. P. Shildneck, and L. A. March, "Equivalent Reactance of Synchronous Machines," AIEE Transactions, Vol. 53, January 1934, pp. 124-132. 25. S. B. Crary, A. H. Lauder, and D. R. Shoults, "Pull-In Characteristics of Synchronous Motors," AIEE Transactions, Vol. 54, 1935, pp. 1385-1395. 26. Gabriel Kron, "The Application of Tensors to the Analysis of Rotating Electrical Machinery," General Electric Review, a serial, April 1935 to December 1937 (Reprinted in book form, 1938). 27. P. S. Shdanov, "Asynchronous Behavior of Electrical Systems," Elek., November 1936, No. 21, pp. 17-26. 28. A. Berger, "Synchronous Machinery," Leningrad, ONTI, 1936 (book). 29. S. B. Crary, "Two-Reaction Theory of Synchronous Machines," Electrical Engineering, Vol. 56, January 1937, pp. 27-31 and p. 36. 30. C. Concordia and H. Poritsky, "Synchronous Machine with Solid Cylindrical Rotor," Electrical Engineering, Vol. 56, January 1937, pp. 49-58 and p. 179. 31. J. W. Butler and C. Concordia, "Analysis of Series Capacitor Application Problems," AIEE Transactions, Vol. 56, 1937, pp. 975-988. 32. C. Concordia, "Two-Reaction Theory of Synchronous Machines with Any Balanced Terminal Impedance," Electrical Engineering, Vol. 56, September 1937, pp. 1124-1127.

BIBLIOGRAPHY 219 33. J. Fallou, "Asynchronous Behavior and Spontaneous Resumption of Synchronism in Interconnected Systems," Revue genirale de Velectricite, Vol. 41, No. 15, April 10, 1937, pp. 451-465. ^^34. A. R. Miller and W. S. Weil, Jr., "Alternator Short-Circuit Currents under Unsymmetrical Terminal Conditions," AIEE Transactions, Vol. 57, October 1937, pp. 1268-1276. 35. A. R. Miller and W. S. Weil, Jr., "Operational Solution of A-C Machines," Electrical Engineering, Vol. 55, November 1937, pp. 1191-1200. 36. B. R. Prentice, "Fundamental Concepts of Synchronous-Machine Reactances," AIEE Transactions (Supplement), Vol. 56, 1937, pp. 1-21. 37. Edith Clarke, C. N. Weygandt, and C. Concordia, "Overvoltages Caused by Unbalanced Short CircuitsEffect of Amortisseur Windings," AIEE Transactions, Vol. 57, August 1938, pp. 453-466. 38. C. Concordia, S. B. Crary, and J. M. Lyons, "Stability Characteristics of Turbine Generators," AIEE Transactions, Vol. 57, 1938, pp. 732-744. 39. E. C. Whitney and H. E. Criner, "Determination of Short-Circuit Torques in Turbine Generators by Test," AIEE Transactions, Vol. 59, 1940, pp. 885-889. 40. M. M. Liwschitz, "Positive and Negative Damping in Synchronous Machines," AIEE Transactions, Vol. 60, 1941, pp. 210-213. 41. C. Concordia and G. K. Carter, "Negative Damping of Electrical Machinery," AIEE Transactions, Vol. 60, 1941, pp. 116-118. 42. C Concordia, S. B. Crary, and Gabriel Kron, "The Doubly Fed Machine," AIEE Transactions, Vol. 61, 1942, pp. 286-289. 43. Gabriel Kron, "Equivalent Circuits for the Hunting of Electrical Machinery," AIEE Transactions, Vol. 61, 1942, pp. 290-296. 44. Gabriel Kron, A Short Course in Tensor Analysis for Electrical Engineers, John Wiley & Sons, 1942. 45. C. Concordia, "Steady-State Stability of Synchronous Machines as Affected by Voltage Regulator Characteristics," AIEE Transactions, Vol. 63, 1944, pp. 215-220. 46. C. Concordia, S. B. Crary, C. E. Kilbourne, and C. N. Weygandt, Jr., "Synchronous Starting of Generator and Motor," AIEE Transactions, Vol. 64, 1945, pp. 629-634. 47. Gabriel Kron and C. Concordia, "Damping and Synchronizing Torques of Power Selsyns," AIEE Transactions, Vol. 64, 1945, pp. 366-371. 48. C. Concordia and M. Temoshok, "Resynchronizing of Generators," AIEE Transactions, Vol. 66, 1947, pp. 1512-1518. 49. Gabriel Kron, "Steady-State Equivalent Circuits of Synchronous and Induction Machines," AIEE Transactions, Vol. 67, 1948, pp. 175-181. 50. C. Concordia, "Synchronous-Machine Damping and Synchronizing Torques," AIEE technical paper 51-122, AIEE Transactions, Vol. 70, 1951.

_EMPTY_

INDEX Air-gap flux, 12 a, 0, 0 components, 77, 78 armature voltage equations, 80 Amortisseur, 4, 6 Amortisseur bars, 6 Armature inductance, 10 Armature transients, neglect of, 68, 186 Armature voltage equations, o, b, c, 8 d, q, 0, 17 a, /3, 0, 80 Asynchronous operation, 165 Axes, direct and quadrature, 6 Ceiling, exciter, 195 Components, a, 0, 0, 77, 78 d, q, 0, 14-17 /, b, 0, 166 1, 2, 0 (symmetrical), 103 Constant flux linkage, 34, 57, 76 Current-density wave, 29 Damper winding, 4 Decrement factors, double-line-toground fault, 108 short-circuit torque, 142 single-phase faults, 90 three-phase fault, 63 Degrees, electrical, 1 Direct axis, 6 amortisseur, 9 subtransient reactance, 57 synchronous reactance, 16 tests for, 23, 24 transient reactance, 35 Double-line-to-ground fault, 100 current, a, 0, 101 6, c, 111 open-phase voltage, 113 torque, 143, 161 Electrical degrees, 1 Energy, stored electromagnetic, 212 Equivalent static circuit, 19, 126, 171 Exciter ceiling, 195 Exciter response, 197 Faults, see Short circuit Field current, double-line-to-ground short circuit, 112 in terms of armature voltage, 24 on application of load, 201 single-phase short circuit, 95 three-phase short circuit, 72 Field excitation, change of, 191 effect on starting torque, 180 in terms of armature voltage, 24 Field flux linkage, 9 constant, 34 in terms of armature voltage, 35 Flux, air gap, 12 Flux-density wave, 29 Flux linkage, 9, 10 constant, 34, 57, 76 field, 34 trapped, 60 Flux-linkage factors, 135 for l-l fault, 136-138 Flux linkages as a function of armature current, in terms of id, iq, io, 18 in terms of symmetrical components, 150 Forward- and backward-rotating quantities, 166 Fourier series, 205 Harmonic components, l-l fault, current, 85, 94 flux linkage, 86 open-phase voltage, 87 torque, 139 l-n fault, current, 86, 94 torque, 146 ll-n fault, current, 206 Impedance, operational, 18, 56 standstill, 125 221

222 INDEX Inductance, armature, 10 rotor, 12 Inductance relations, 10 Infinite bus, 26 Leakage reactance, 3 Line-to-ground fault, 81 current, 86 open-phase voltages, 83 torque, 143, 160 Line-to-line fault, current, 76 open-phase voltage, 87 torque, 139, 160 Maximum power, 40 Mmf, rotor, 91 Negative-phase-sequence reactance, 87, 89 Reactance, leakage, 3 negative-phase sequence, 87, 89, 102, 107 subtransient, direct axis, 57 quadrature axis, 57 synchronous, direct axis, 16 quadrature axis, 16 transient, direct axis, 35 zero-phase sequence, 16 Reactive volt-amperes, 44 versus angle, 45 Reciprocity of mutual inductances, 19 Reluctance torque, 30 Response of exciter, 197 Rotor-circuit currents with three-phase faults, 55 Rotor circuits, 6 Rotor inductances, 12 Rotor mmf's, 91 Saliency, 6 effect on steady-state torque, 39 subtransient, 58 transient, 189 Saturation, 196 Sequential faults, 114 Short circuit, decrement factors for, 90, 142 for double-line-to-ground fault, 102, 107 Open-phase voltage, for double-line-toground faults, 113 for line-to-line faults, 87 Operational impedances, 18, 56 Permeance, 11 Per-unit quantities, 20 Per-unit time, 22 Power, input, 27 limit, 40 maximum, 40 output, 25, 36 Power-angle characteristics, 27, 37 for two machines, 46-48 with constant flux linkages, 38 Problems, for Chapter 2, 31 for Chapter 3, 52 for Chapter 4, 75 for Chapter 5, 98 for Chapter 6, 118 for Chapter 7, 164 for Chapter 8, 184 for Chapter 9, 203 Quadrature axis, 6 amortisseur, 9 subtransient reactance, 57 synchronous reactance, 16 double line to ground, 100 field current, 72, 95 line to line, 76 line to neutral, 81 sequential, 114 test for Xd, 24 three phase, 54 from load, 58 oscillograms, 66 steady-state current, 59 transient current, 65 with armature resistance, 67 time constants, 63, 142 torques, 119 Short-circuit current, double line to ground, 111, 112 fundamental-frequency components of id, iq, 157 line to ground, 94 line to line, 94 three phase, 58, 66, 71 Sinusoidally distributed winding, 7

INDEX 223 Slip, 165 Slip test for synchronous reactance, 23 Slots, stator, 8 Speed change, 130 Stability, 40 Starting torque, 165 see also Torque Steady-state performance, 32, 165 Steady-state short-circuit current, 59 Subtransient reactances, 57 Subtransient time constant for threephase fault, 62 Symmetric-rotor machine, 151 Symmetrical components, 103 mathematical definition, 104 of torque, 149 physical meaning, 103, 106 relation to a, /3, 0 components, 105 short-circuit currents, 152 Synchronism, 3 Synchronizing, 58 Synchronous reactance, direct and quadrature axis, 16 short-circuit test for, 24 slip test for, 23 Tests, for synchronous reactance, 24 short circuit, 66 Three-phase short circuit, current, 66 field current, 72 torque, 145, 161 with armature resistance, 67 Time constant, field load, 69, 188 field open circuit, 64 rotor open circuit, 64 rotor short circuit, 124 short circuit, 142-144 double line to ground, armature, 110 rotor, 108 line to line, armature, 90 rotor, 93 line to neutral, armature, 90 rotor, 95 three phase, 60, 63, 64 Torque, derivation of equations for, 28, 212 harmonic components, of l-l, 139 of l-n, 146 in terms of complex quantities, 148 in terms of ia, i&, ic, 215 Torque, in terms of id, iq, 28 in terms of ia, ip, 147 in terms of symmetrical components, 149-152 reluctance, 30 short circuit, 119 approximate equations for, 160 examples of, 162 general formulas for, 142 line to line, 134 examples of, 140-141 three phase, 119 components of, 127 effect of rotor resistance, 123 effect of stator resistance, 121 maximum unidirectional, 122, 133 neglecting resistance, 120 unidirectional, 147 starting, 165 approximate equation for, 160 at half speed, 174 average, 170, 177 comparison of exact and approximate, 175 equivalent circuit for, 171 in terms of id, iq, 177 oscillating component of, 171, 183 with field excitation, 180 Transformation equations, to d, q, 0 axes, 17 current, 14 flux linkage, 15 voltage, 16 Transformations, a, /3, 0 axes, 77 d, q, 0 axes, 14-17 /, b, 0 axes, 166 1,2,0 (symmetrical components) axes, 104 Transients, neglect of armature, 68, 186 Trapped flux, 60 Turn ratio, 21 Vector diagram, 32 construction, 33 with resistance load, 70 Voltage, armature, 8, 17 back of transient reactance, 35 open phase, for l-l fault, 84 for ll-n fault, 113

224 INDEX Voltage dip, 185 estimating chart, 199 field-current change, 201 initial, 189 minimum voltage, 193 with initial load, 200 Voltage equations, armature, 8, 17, 80 Voltage regulator, 185, 191 c Voltage rise on application of load, 189 Volt-amperes, reactive, 44 Winding, amortisseur (or damper), 4 sinusoidally distributed, 7 Zero-phase-sequence reactance, 16 test for, 25 152 69 220 AA A 30

_EMPTY_

_EMPTY_

_EMPTY_

~ a J UNIVERSITY OF MICHIGAN 3 9015 00201 9779

, i

H I

You might also like