You are on page 1of 631
EMTP THEORY BOOK Second Edition Hermann W. Dommel Microtran Power System Analysis Corporation Vancouver, British Columbia May, 1992. Latest update: April 1996 Copyright © 1996 by H. W. Dommel All rights reserved FOREWORD TO THE SECOND EDITION This second edition includes a few improvements and the correction of errors which have been found by readers and by myself. I am grateful to engineers of the System Planning Department of Taiwan Power Company, who discovered many of the errors when I used the manual for the first time during a short course in 1987. Errors were also reported to me by A.B. Araujo, S. Carneiro Jr., J. Lin, J.R. Marti and L. Marti. Section 4 was modified to explain the derivation of the equivalent geometric mean radius of bundle conductors and of stranded conductors, with the details put into a new Appendix VIII. G. Calzolari and C. Saldafia went carefully through Section 6 (Transformers) during my stay in Itajubé, Brazil, in 1989. I owe most of the improvements in that Section to their suggestions. Section 8 (Three-Phase Synchronous Machine) and Section 9 (Universal Machine) have been improved based on suggestions of E.F. Fuchs. Discussions with F.L. Alvarado and C. Cafiizares in 1988 made me aware of subtleties in the data conversion of synchronous machine parameters. This led to revisions of Section 8.2 and of Appendix VI, with the help of my wife 1.1. Dommel. P. Barret, B. Blengino and D. Souque of Electricité de France pointed out to me that the correction for saturation in the steady-state operation of the synchronous machine is direct (non-iterative) if saturation is applied to the total flux. Section 8.6.2 has been modified accordingly. In addition to the two major versions of the EMTP mentioned in this manual (BPA EMTP and UBC EMTP), other versions have appeared in the meantime. There are now three major version: The DCG/EPRI EMTP is the product of the EMTP Development Coordination Group (DCG), which was formed in 1982, and joined by the Electric Power Research Institute of the U.S.A. in 1984. The ATP (Alternative Transients Program) is being developed through the EMTP Centre of the Catholic University in Leuven, Belgium. Finally, an expanded version of the UBC code is now available for IBM compatible personal computers through Microtran Power System Analysis Corporation (a UBC spin-off company), under the name MicroTran®. It includes, among other features, frequency-dependent transmission lines, thyristor ro a models and elimination of numerical oscillations with the critical damping adjustment procedure. It would have been a major undertaking to include all the improvements of these EMTP versions in this second edition, even if all the information were freely available. Maybe it can be attempted in a future edition. In the meantime, readers interested in the following major changes should consult the literature quoted in parenthesis: 1. Critical damping adjustment scheme for suppression of numerical oscillations (J.R. Marti and J. Lin, "Suppression of numerical oscillations in the EMTP*, IEEE Trans. Power Systems, vol.4, pp. 739-749, May 1989. J. Lin and J.R. Marti, “Implementation of the CDA procedure in the EMTP", IEEE Trans.Power Systems, vol. 5, pp. 394-402, May 1990). 2. Corona on transmission lines (P.S. Maruvada, D.H. Nguyen, H. Hamadani-Zadeh, "Studies on modeling corona attenuation of dynamic overvoltages", IEEE Trans. Power Delivery, vol. 4, pp. 1441-1449, April 1989. S. Carneiro Jr. and J.R. Marti, “Evaluation of corona and line models in electromagnetic transients calculations", IEEE Trans. Power Delivery, vol. 6, pp. 334-342, Jan. 1991). 3. Circuit breakers (V. Phaniraj and A.G. Phadke, "Modelling of circuit breakers in the electromagnetic transients program’, IEEE Trans. Power Systems, vol 3, pp. 799-805, May 1988). Last, but not least, I would like to thank Mae Marti for completely retyping Appendix VI, and L.M. Spencer for retyping most of the other changes. H.W. Dommel Vancouver, Canada, November, 1991 Remarks about latest update of April 1996 Dr. D. H. Boteler, Geological Survey of Canada, pointed out a typing error in Carson's equation for the earth return impedance. Mr. H. Hamadanizadeh found some errors in cable impedance formulas. Prof. A. E. A. Araujo, Federal University of Minas Gerais, Belo Horizonte, Brazil, and Mr. 0. Trad, National University of San Juan, Argentina, discovered errors in various places. I am grateful to these readers for bringing errors to my attention; they have all been corrected in this update of April 1996. 444 ‘ACKNOWLEDGEMENT Without the encouragement from Dr. W.S. Meyer, Dr. Teu-huei Liu, and Dr. J. Vithayathil of Bonneville Power Administration, this EMTP REFERENCE MANUAL would never have been written. They suggested that such a manual be compiled in the Summer of 1981, and they have helped we in the following years to put it together. While I wrote large parts of this manual myself, and take responsibility for its entire contents, I could not have done it without the valuable assistance of my co-authors L. Marti for Section 5.9 (Cable Models in the EMTP), V. Brandwajn for Section 8 (Three-Phase Synchronous Machine), W.K. Lauw for Section 9 (Universal Machine) and S. Bhattacharya for Section 13 (Transient Analysis of Control Systems). J. Mechenbier contributed the synchronous machine test case of Table 8.2. Other contributions are acknowledged in the text. Draft copies of the manual were submitted for review to some specialists, and their comments have been taken into account. The reviews of Mr. Robert M. Hasiber and Mrs. Li Guang Qi were particularly useful. ‘Typing the manual with the many complicated equations was no doubt a difficult assignment. I am very grateful to the following staff members of the Department of Electrical Engineering at the University of British Columbia for their excellent work: Kathy Brindamour, Michele Depuit, Gail Schmidt and Charlotte Stevenson. Dr. L.M. Wedepohl (Acting Head in 1981) and Dr. K-D. Srivastava (Head of the Department) kindly permitted the typing to be done in the Department. I also wish to thank the authors and publishers mentioned in figure captions for their permission to reprint copyrighted material. I thank my daughter Monica K. Dommel for inserting the figures into the text. Last but not least, I want to acknowledge the assistance of my wife Irmgard I. Donel. She did most of the calculations which were needed to fill gaps in the theory and ran many EMTP cases needed to illustrate the theory. She also drew most of the figures and looked after the layouts and proofreading. Hermann W. Doumel Vancouver, Canada, August 1986 iv TABLE OF CONTENTS Preface Acknowledgement Table of Contents ae a INTRODUCTION TO THE SOLUTION METHOD USED IN THE ENTP LINEAR, UNCOUPLED LUMPED ELEMENTS 2.1 Resistance R 2.1.1 Error analysis 2.1.2 Example for network with resistances 2.2 Self Inductance L 2.2.1 Error analysis 2.2.2 Damping of "numerical oscillations" with parallel resistance 2.2.3 Physical reasons for parallel resistance 2.2.4 Example for network with inductances 2.3 Capacitance c 2.3.1 Error analysis 2.3.2 Damping of "numerical oscillations" with series resistance 2.3.3 Physical reasons for series resistance 2.3.4 Example for network with capacitances 2.4 Series Connection of R, L, C 2.5 Single-phase Nominal II-Circuit LINEAR, COUPLED LUMPED ELEMENTS 3.1 Coupled Resistances [R] 3.1.1 Error analysis 3.1.2 Insertion of coupled branches into nodal equations 3.1.3 Example of coupled resistances 3.2 Coupled Inductances [L] 3.2.1 Error analysis Page ai iii iv a 2-21 ae 2 2-4 2-5 2- 9 2- 14 2- 20 2- 24 2- 26 2- 28 2- 30 2- 31 2- 33 1 3-1 3- 3 3-3 3- 5 3- 8 3- 10 3.2.2 Damping of "numerical oscillations" with coupled parallel resistances ete 3.2.3 Physical reasons for coupled parallel resistances 3-1 3.2.4 Example for network with coupled inductances 3- 11 3.3 Coupled Capacitances [C] 3-13 3.3.1 Error analysis 3- 16 3.3.2 Damping of "numerical oscillations" with series resistances 3-16 3.3.3 Physical reasons for coupled series resistances 3-17 3.3.4 Example for network with coupled capacitances 3-17 3.4 M-Phase Nominal II-circuit 3- 20 3+4.1 Series connection of [R] and (1]_, 3- 20 3.4.2 Series connection of [R] and [L] 3-21 OVERHEAD TRANSMISSION LINES Bs 4.1 Line Parameters a1 4.1.1 Line parameters for individual conductors 4-2 4.1.1.1 series impedance matrix 4-3 4.1.1.2 shunt capacitance matrix 4-15 4.1.1.3 insulated overhead conductors 4-18 4.1.2 Line parameters for equivalent phase conductors 4-18 4.1.2.1 elimination of ground wires 4-18 4.1.2.2 bundling of conductors 4-19 4.1.2.3 reduced matrices for equivalent phase conductors 4-23 4.1.2.4 nominal I1-circuit for equivalent phase conductors 4-25 4.1.2.5 continuous and segmented ground wires 4-28 4.1.3 Positive and zero sequence parameters of balanced lines 4-33 4.1.3.1 positive and zero sequence parameters of single-circuit three-phase lines 4-35 4.1.3.2 positive and zero sequence parameters of balanced M-phase lines 4.1.3.3 two identical three-phase lines with Zero sequence coupling only 4-44 4.1.4 Symmetrical components 4.1.5 Modal parameters 4.1.5.1 line equations in modal domain 4.1.5.2 lossless high frequency approximation 4.1.5.3 approximate transformation matrices for transient solutions 4.2 Line Models in the EMTP 4.2.1 Ac steady-state solutions 4.2.1.1 nominal M-phase II-circuit 4.2.1.2 equivalent II-circuit for single- phase lines 4.2.1.3 equivalent M-phase II-circuit 4.2.2 Transient solutions 4.2.2.1 nominal II-circuits 4.2.2.2 single-phase lossless line with constant L' and c! 4.2.2.3 M-phase lossless line with constant L' and c! 4.2.2.4 single and M-phase distortionless lines with constant parameters 4.2.2.5 single and M-phase lines with lumped resistances 4.2.2.6 single and M-phase lines with frequency-dependent parameters UNDERGROUND CABLES 5.1 Single-core Cables 5.1.1 Series impedances 5.1.2 Shunt admittances 5.2 Parallel Single-core Cables 5.3 Earth-Return Impedances 5.3.1 Buried conductors in semi-infinite earth 2 Buried conductors in infinite earth 3 Overhead conductors 4 Mutual impedance between overhead conductor and buried conductor 5. 5. 5. wou 5.4 Pipe-Type Cable 5.4.1 Infinite pipe thickness (no earth return) 5.4.2 Finite pipe thickness with earth return 5.5 Bundling of Conductors and Elimination of Grounded Conductors a re 50 54 87 66 67 67 69 2 72 76 83 89 90 93 21 23 27 28 vit Buried Pipelines Partial Conductor and Finite Element Methods 5.7.1 Subdivision into partial conductors 5.7.2 Finite element methods Modal Parameters Cable Models in the EMTP Ac steady-state solutions Transient solutions 5.9.2.1 short cables 5.9.2.2 single-phase cables 5.9.2.3 polyphase cables 5.9.1 5.9.2 TRANSFORMERS 6.1 6.2 Transformers as Part of Thevenin Equivalent Circuits Inductance Matrix Representation of Single-Phase Two- and Three-Winding Transformers 6.2.1 Two-winding transformers 6.2.2 Ill-conditioning of inductance matrix 6.2.3 Three-winding transformers Inverse Inductance Matrix Representation of Single- Phase Two- and Three-Winding Transformers 6.3.1 Two-winding transformers 6.3.2 Three-winding transformers Matrix Representation of Single-Phase N-Coil Transformers Matrix Representation of Three-Phase N-Coil Transformers a 6.5.1 Procedure for obtaining [R] and [L] 6.5.2 Modification of zero-sequence data for delta connections Exciting Current 6.6.1 Linear (unsaturated) exciting current 6.6.1.1 single-phase transformers 6.6.1.2 three-phase transformers 6.6.2 Saturation effects 5- 30 5- 36 5- 36 5- 38 5- 40 5- 41 5- 42 5- 42 5- 42 5- 43 5-44 6- 1 6- 2 6- 4 6- 4 6 7 6- 8 6- 9 6- 9 6- 10 6- 19 6- 21 6- 24 6- 25 6- 26 viii 1 single-phase transformers 2 three-phase transformers 6.6.3 Hysteresis and eddy current losses 6.6.4 Residual flux 6.7 Autotransforners 6.8 Ideal Transformer 6.9 Floating Delta Connections 6.10 Description of Support Routines and saturable ‘Transformer Component 6.10.1 Support routine XFORMER 6.10.2 Support routine BCTRAN 6.10.3 Support routine TRELEG 6.10.4 Support routine CONVERT 6.10.5 Saturable transformer component 6.11 Frequency-Dependent Transformer Models SIMPLE VOLTAGE AND CURRENT SOURCES 7.1 Connection of Sources to Nodes 7.2 Current Sources Between Two Nodes 7.3 Voltage Sources Between Two Nodes 7.4 More Than One Source on Same Node 7.5 Built-in Simple Source Functions 7.6 Current-controlled de Voltage Source 7.6.1 Steady-state solution 7.6.2 Transient solution THREE-PHASE SYNCHRONOUS MACHINE 8.1 Basic Equations for Electrical Part 8.2 Determination of Electrical Parameters 8.3 Basic Equations for Mechanical Part 6- 35 6- 36 6- 43 6- 49 6- 50 6- 51 6- 52 6- 52 6- 54 6- 56 T- 1 1 2 2 4 5 I- 9 7-11 7-12 8-1 B- 3 8- 10 8- 17 9. 8.4 steady-state Representation and Initial Conditions 8.4.1 Initialization with positive sequence values 8.4.2 Initialization with negative sequence values 8.4.3 Initialization with zero sequence values 8.4.4 Initialization of the mechanical part Transient Solution 8.5.1 Brief outline of solution method 8.5.2 Transient solution of the electrical part 8.5.3 Transient solution of the mechanical part 8.5.4 Prediction and correction schemes 8.5.4.1 prediction of a ands 8.5.4.2 averaging of d- and q-axis companion resistances 3 prediction of i,, i. 4 prediction of sfeedYvoltages 5 8.5.4 8.5.4 8.5.4.5 iteration schemes Saturation 8.6.1 Basic assumptions 8.6.2 Saturation in steady-state operation 8.6.3 Saturation under transient conditions 8.6.4 Implementation in the EMTP 8.6.4.1 steady-state initialization 8.6.4.2 transient solution 8.6.5 Saturation effects with Canay's characteristic reactance UNIVERSAL MACHINE 9.1 9.2 9.3 9.4 9.5 Basic Equations for Electrical Part Determination of Electrical Parameters Transformation to Phase Quantities Mechanical Part Steady-State Representation and Initial a 8 a a a a- a B- 8 8 22 26 29 35 35 38 38 42 49 50 50 54 54 57 59 60 60 65 68 70 2 72 73 10 13 Conditions 9-15 Three-phase synchronous machine 9- 16 Two-phase synchronous machine 9-17 Single-phase synchronous machine 9-17 De machines 9- 18 Three-phase induction machine 9- 18 Two-phase induction machine 9- 22 Single-phase induction machine 9- 22 Doubly-fed induction machine 9- 23 9.6 Transient Solution with Compensation Method 9- 24 9.7 Transient Solution with Armature Flux Prediction 9- 28 9.8 Saturation 9- 32 10. SWITCHES lo- 2 10.1 Basic Switch Types lo- 2 10.1.1 Time-controlled switch lo- 2 10.1.2 Gap switch lo- 4 10.1.3 Diode switch lo- 6 10.1.4 Thyristor switch (TACS controlled) lo- 6 10.1.5 Measuring switch lo- 7 10.2 Statistical Distribution of switching Overvoltages 1o- 7 10.2.1 Statistics switch 10- 8 10.2.2 Systematic switch 10- 11 10.3 Solution Methods for Networks with Switches 10- 12 10.3.1 Network reduction to switch nodes 10- 13 10.3.2 Complete re-triangularization 10- 15 10.3.3 Switch closing 10- 15 10.3.4 Switch opening 10- 16 10.4 Arc Phenomena in Circuit Breakers 10-17 11. SURGE ARRESTERS AND PROTECTIVE GAPS ui- 1 11.1 Protective Gaps ui- 1 11.2 surge Arresters li- 2 11.2.1 Silicon-carbide surge arrester ll- 2 11.2.2 Metal-oxide surge arrester ll- 5 xi 12. SOLUTION METHODS IN THE EMTP 12.1 12.2 12.3 12.4 Inclusion of Nonlinear Elements 12.1.1 Current-source representation with time lag At 12.1.2 Compensation method 12.1.2.1 one nonlinear element 12.1.2.2 two or more nonlinear elements separated by travel time 12.1.2.3 two or more connected nonlinear elements nonlinear inductance Newton-Raphson method numerical problems 12.1.3 Piecewise linear representation 12.1.3.1 piecewise linear inductance 12.1.3.2 piecewise linear resistance 12.1.3.3 numerical problems Load Flow option 12.2.1 Rasmussen's load flow method 12.2.2 Current source iteration method Steady-State Solutions without Harmonics 12.3.1 Thevenin equivalent circuits 12.3.2 Frequency scan 12.3.3 Different frequencies in disconnected parts Steady-State Solution with Harmonics 12.4.1 Harmonics from linear ac steady-state solution followed by transient simulation 12.4.2 Harmonics from steady-state solutions 12.4,.2.1 "power flow" iterations 12.4,.2.2 "distortion" iterations 12.4.3 Discrepancies between harmonics in steady-state and transient solution 12.4.4 Ferroresonance 13. TRANSIENT ANALYSIS OF CONTROL SYSTEMS (TACS) 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12- 12 a2- 12- 12- 12- 13- 10 12 13 13 17 17 20 22 29 30 33 35 36 37 37 39 41 43 45 xit 13.1 Interface between TACS and the Electric Network 13.2 Transfer Function Block with Summer 13.3 Limiters 13.3.1 Windup limiter 13.3.2 Non-windup limiter 13.4 Limiter Implementation with Possible Time Delay 13.5 Signal Sources 13.6 Special Devices 13.6.1 FORTRAN-defined special devices 13.6.2 Built-in special devices 13.7 Initial Conditions APPENDIX I NUMERICAL SOLUTION OF ORDINARY DIFFERENTIAL EQUATIONS I.1 Closed-Form Solution I.2 Taylor Series Approximation of Transition Matrix I.3 Rational Approximation of Transition Matrix I.4 Trapezoidal Rule of Integration I.5 Runge-Kutta Methods I.6 Predictor-Corrector Methods I.7 Deferred Approach to the Limit (Richardson Extra- polation and Romberg Integration) I.8 Numerical Stability and Implicit Integration I.9 Backward Euler Method APPENDIX II RE-INITIALIZATION AT INSTANTS OF DISCONTINUITY APPENDIX III SOLUTION OF LINEAR EQUATIONS, MATRIX REDUCTION AND INVERSION, SPARSTTY III.1 Gauss Elimination III.2 Gauss-Jordan Elimination by Diagonalization III.3 Subroutines REDUCT and CXRED for Matrix Inversion, Reduction and Solution of Equations with Symmetric Matrices III.4 Gauss Elimination with Sparsity Techniques III.4.1 Basic idea III.4.2 Row-by-row elimination with static storage 31.4.3 Working row 13- 2 13- 5 13- 10 13- 10 13- 13 13- 15 13-17 13- 18 13- 19 13- 19 13- 20 I-. I-1 I-7 I-38 I-9 I-12 I-15 I-18 I-19 1-23 TI-1 III- 1 III- 3 III- 8 III-10 III-12 III-12 III-15 III-18 xiii III.4.4 Row pointer/column index storage scheme TII.4.5 Special techniques for symmetric matrices APPENDIX IV ACTUAL VALUES VERSUS PER-UNIT QUANTITIES IV.1 Per-Unit Quantities IV.2 Conversion from One Base to Another IV.3 Actual Values Referred to One Side of Transformer IV.4 Advantages of Actual Values IV.5 Per-Unit Voltages with Actual Impedances APPENDIX V RECURSIVE CONVOLUTION APPENDIX VI TRANSIENT AND SUBTRANSIENT PARAMETERS OF SYNCHRONOUS MACHINES VI.1 Transient Parameters with only One Winding on the Pield structure VI.2 Subtransient and Transient Time Constants with Two Windings on the Field Structure VI.3 Subtransient and Transient Reactances with Two Windings on the Field Structure VI.4 Canay's Data Conversion VI.5 Negative Sequence Impedance APPENDIX VII INTERNAL IMPEDANCE OF STRANDED CONDUCTORS APPENDIX VIII GEOMETRIC MEAN RADIUS OF BUNDLE CONDUCTORS AND STRANDED CONDUCTORS VIII.1 Bundle Conductors VIII.2 Stranded Conductors REFERENCES III-19 III-21 viI- 1 vI- 1 vI- 2 VI- 5 VI- 9 VI-12 VII- 1 VIII- 1 VIII- 1 VIII- 4 1-1 INTRODUCTION 10 THE SOLUTION METHOD USED IN THE EMTP ‘This manual discusses by and large only those solution methods which are used in the EMTP, It is therefore not a book on the complete theory of solu~ tion methods for the digital simulation of electromagnetic transient pheno mena. ‘The developers of the EMIP chose methods which they felt are best suited for a general-purpose program, such as the EMTP, and it is these methods which are discussed here. For analyzing specific problems, other methods may well be competitive, or even better. For example, Fourier trans- formation methods may be preferable for studying wave distortion and attenua~ tion along a line in cases where the time span of the study is so short that reflected waves have not yet come back from the remote end. ‘The ENTP has been specifically developed for power system problems, but some of the methods have applications in electronic circuit analysis as well. While the developers of the EMTP have to some extent been aware of the methods used in electronic circuit analysis programs, such as TRAC or ECAP, the reverse may not be true. A survey of electronic analysis programs pub- lished as recently as 1976 [22] does not mention the EMTP even once. Computer technology is changing very fast, and new advances may well make this manual obsolete by the time it is finished. Also, better numerical solution methods may appear as well, and replace those presently used in the EMTP. Both prospects have been discouraging for the writer of this manual; what has kept him going is the hope that those who will be developing better programs and who will use improved computer hardware will find some useful information in the description of what exists today. Digital computers cannot simulate transient phenomena continuously, but only at discrete intervals of time (step size At). This leads to truncation errors which may accumulate from step to step and cause divergence from the true solution. Most methods used in the EMTP are numerically stable and avoid this type of error build-up. ‘The EMIP can solve any network which consists of interconnections of resistances, inductances, capacitances, single and multiphase r-circuits, istributed-parameter lines, and certain other elements. To keep the explanations in this introduction simple, only single-phase network elements 12 will be considered and the more complex multiphase network elements as well as other complications will be discussed later. Fig. 1-1 shows the details of a larger network just for the region around node 1. Suppose that voltages and currents have already been computed at time instants 0, At, 2At, etc., up to t-At, and that the solution mst now be found at instant t. At any instant of time, the sun of the currents flowing away from node 1 through the branches must be equal to the injected current 4): type) + 1,5(t) + y(t) + Ly5(e) = 4, Ct). qa. lossless distributed para~ meter line with Z, T Fig. 1-1 - Details of a larger network around node no. 1 Node voltages are used as state variables in the EMTP. It is therefore necessary to express the branch currents, i, etc., as functions of the node voltages. For the resistance, 4190) = x {vy Ce) ~ vp Ce} a2 For the inductance, a simple relationship is obtained by replacing the differential equation veils with a central difference equation v(t) + v(t = at) |, A(t) = 1(t - ae) 2G) + ye a oe) 2 Ate) en oe) at This can be rewritten, for the case of Fig. 1.1, as 4 g(t) = GE fvy(e) - vy(e)} + hist, 4(e-ae), (3a) with hist ,, known from the values of the preceding time step, hist y(t-st) = Ap y(t-ar) +E fyy(e-ae) - vy(e-ae)}. (1.30) The derivation for the branch equation of the capacitance is analogous, and leads to 44400) = 22 fy (ey ~ vy(e)} + ise, ,(e-ae), (4a) with hist , again known from values of the preceding time step, 2c hist ,(e-st) = ty ,(e-se) ~ ZB fu (tout) - v,Cemae) J. (14d) Readers fresh out of University, or engineers who have read one or one too many textbooks on electric circuits and networks, may have been misled to believe that Laplace transform techniques should be used for solving systems with lumped elements. Laplace transform techniques are only useful for “hand solutions” of rather small networks, and more or less useless for computer solutions of problems of the size typically analyzed with the EMTP. Since even new textbooks perpetuate the myth of the usefulness of Laplace trans~ forms, Appendix I has been added for the mathematically-minded reader to summarize numerical solution methods for linear, ordinary differential equa~ tions. For the transmission line between nodes 1 and 5, losses shall be ignored in this introduction. Then the wave equations ~ He at x 3 Be oe Oo where 1-4 L', C! = inductance and capacitance per unit length”, x = distance from sending end, have the well-known solution due to d'Alembert: 1 = F(x ~ ct) - f(x + ct) v= 2F( x ~ ct) + Zf(x + ct), (1.5a) with F(x = ct) ] . Fx 5 €2} | = functions of the composite expressions x - ct and x + ct, Z = surge impedance Z = YE"/C™ (constant), ¢ = velocity of wave propagation (constant). If the current in Eq. (1.5a) {e multiplied by Z and added to the vol~ tage, then vt Zi = 22P(x - ct). (1.5b) Note that the composite expression v + Zi does not change 1f x - ct does not change. Inagine a ficticious observer travelling on the line with wave velo- city c. The distance travelled by this observer is x = x, + ct (x, = locaé tion of starting point), or x ~ ct is constant. If x ~ ct is indeed con- stant, then the value of v + Zi seen by the observer must also renain con stant. With travel time _ Hine length © an observer leaving node 5 at time t ~ t will see the value of v.(t - 1) + Zig y(t - 1), and upon arrival at node 1 (after the elapse of travel time 1), Will see the value of v(t) ~ Zi y(t) (negative sign because 11, has opposite direction of 1,,). But since this value seen by the observer must remain constant, both of these values must be equal, giving, after rewriting, Ays(t) =F vy(t) + hist, (t - 0, (16a) iste where the term hist, is again known from previously computed values, “the prime is used on L', C’ to distinguish these distributed parameters ng from lumped parameters LC. hist g(t-1) = - P ve(t-r) - 4g (t-). (1,60) Example: Let At = 100 ys and t= Ims, From equations (1.6) it can be seen that the known "history" of the line must be stored over a time span equal to t, since the values needed in Eq. (1.6b) are those computed 10 time steps earlier. Eq. (1.6) is an exact solution for the lossless line if t is an integer multiple of At; if not, linear interpolation is used and inter- polation errors are incurred, Losses can often be represented with suffi- cient accuracy by inserting lumped resistances in a few places along the line, as described later in Section 4.2.2.5. A more sophisticated treatment of losses, especially with frequency dependent parameters, is discussed in Section 4.2.2.6. If Eq. (1.2), (1.3a), (1.4a) and (1.6a) are inserted into Eq. (1.1), then the node equation for node 1 becomes At , 2c i a At fet D V(t) ~ R v_lt) ~ FF vy (t) - gs 2 By ce = ¢ zx R At 4 a(t) 7 hist, (t-At) - hist, ,(t-At) - hist) ,(t-t), (1.7) which is simply a linear, algebraic equation in unknown voltages, with the right-hand side known from values of preceding time steps. For any type of network with n nodes, a system of n such equations can be formed”, (c] {v(t)] = [4(t)] - thist], (18a) with [G] = n x n symmetric nodal conductance matrix, {v(t)] (i(t)] = vector of n current sources, and vector of n node voltages, (hist] = vector of n known "history" terms. Normally, some nodes have known voltages either because voltage sources are connected to them, or because the node is grounded. In this case Eq. (1.8a) is partitioned into a set A of nodes with unknown voltages, and a set B of nodes with known voltages. The unknown voltages are then found by solving Brackets are used to indicate matrix and vector quantities. [Gqq JIv4#) ] = ByCed] - otst,]- [6,5 ]I¥g(t) ] (1.8) for [v,(t) ]- The actual computation in the EMTP proceeds as follows: Matrices [6,4] and [G,,] are built, and [6,,] is triangularized with ordered elimination and exploitation of sparsity. In each time step, the vector on the right-hand side of Eq. (1-8b) is “assembled” from known history terms, and known current and voltage sources. Then the system of linear equations is solved for [vg(t) ], using the information contained in the triangularized conductance matrix. In this “repeat solution” process, the symmetry of the matrix is exploited in the sense that the same triangularized matrix used for downward operations is also used in the backsubstitution. Before proceeding to the next time step, the history terms hist of Eq. (1.3b), (1.4b) and (1.6b) are then updated for use in future time steps. Originally, the EMTP was written for cases starting from zero initial conditions. In such cases, the history terms hist,3, hist ,, and hist ys in Eq. (1.7) are simply preset to zero, But soon cases arose where the transient simulation had to be started from power frequency (50 or 60 Hz) ac steady-state initial conditions. Originally, such ac steady-state initial conditions were read in “), but this put a heavy burden on the program user, who had to use another steady-state solution program to obtain them. Not only was the data transfer bothersome, but the separate steady-state solution program might also contain network models which could differ more or less from those used in the EMIP, It was therefore decided to incorporate an ac steady-state solution routine directly into the EMIP, which was written by JeW. Walker. The ac steady-state solution shall again be explained for the case of Fig. 11, Using node equations again, Eq. (1.1) now becomes Tettsttyths7 tp a.9) where the currents I are complex phasor quantities |T| +e! now. For the this option is still available in the EMIP, but it has become somewhat of a historic relic and has seldom been used after the addition of a steady- state solution routine, For some types of branches, it may not even work (U1, p. 37e)- 1-7 lumped elements, the branch equations are obvious. For the resistance, 1 1, *2W,- vy), (1.10) for the inductance, 13 77 , - Y5) aay 137 Jar Yy ~ Ys)» and for the capacitance, yy 7 5 (VY, Vy, (1.12) For a line with distributed parameters R', L', G’, C", the exact steady- state solution is Tis a (1.13) T51 ‘series : ; : 2 ‘shunt Y, =o, wine =e (R + jul SEBO) : : cng) z Yonunt, z (G' jue’ ayia! (1.14) 2 1-8 and sometimes equally useful, 1 Yeeries Mes Yohunt, = cosh(yt) * series’ where y is the propagation constant, yt RF oF a) (is) For the lossless case with R' = 0, and G' = 0, Eq. (1-14) simplifies to ceten * tegutt « SIMONE, wh/T'T at L a . tan(y— yITCT) “shunt 72° 8 * Sa? eee Ze 1 . a Yeertes * Z Yshunt ~ COMET) * Yeertes If the value of wt is emall, typically 2 < 100 km at 60 Hz for overhead sinh(x) 444 tanh(x/2) Lines, then the ratios S484) ang £80NG/?) tn eq. (1.14), a8 well as sin(x) sag ee x is usually called the “nominal” x-circuit, tan(x/2) * a ee sanG/?) im Bq. (1-16) all become 1.0. This simplified n-etreuit Zeertes 7 * (RI + Sul") 1 1 ) af wt te onatl. aur Z Yenune ~ Z (6 + $00") With the equivalent circuit of Fig. 1-2, the branch equation for the loss~ less line finally becomes a y, Ts Ceeries 2 Z Yehunt¥1 ~ “serie: : (18) Now, we can again write the node equation for node 1, by inserting Eq. (110), (1.11), (1.12) and (1.18) into Eq. (1.9), : w,- dv l,ly, 1 (et gar * C+ Xoertes* Z Yenune M27 & V2” Jor Y37 $46 V4" Yeertes’5™ 1 a (19) For any type of network with n nodes, a system of n such equations can be formed, 19 (itv) = (21, (1.20) with [Y] = symmetric nodal admittance matrix, with complex elements, [V] = vector of n node voltages (complex phasor values), [1] = vector of n current sources (complex phasor values). Again, Eq. (1-20) is partitioned into a set A of nodes with unknown voltages, and a set B of nodes with known voltages. ‘The unknown voltages are then found by solving the system of linear, algebraic equations [iggllvg) = Ug) - [gg] l¥gl (21) Bringing the term [Y¥,g]{Vg] from the left-hand side in Eq. (1-20) to the right-hand side in Eq. (1-21) is the generalization of converting Thevenin equivalent circuits (voltage vector [V,] behind admittance matrix [¥,g]) into Norton equivalent circuits (current vector [Y,,]{V,] in parallel with admtt- tance matrix [¥,51)+ 2-1 2. LINEAR, UNCOUPLED LUMPED ELEMENTS Linear, uncoupled lumped elements are resistances R, self inductances L, and capacitances C. They usually appear as parts of equivalent circuits, which may represent generators, transformers, short sections of transmission lines, or other components of an electric power system, or they may represent ‘a component by itself. 2.1 Resistance R Resistance elements are used to represent, among other things, (a) closing and opening resistors in circuit breakers, () tower footing resistance (as a crude approximation [8] of a compli- cated, frequency-dependent grounding impedance), (ce) resistance grounding of transformer and generator neutrals, (a) “metering” - resistance in places where currents or branch voltages cannot be obtained in other ways by the EMTP, (e) as parts of equivalent networks, e.g., in parallel with inductances to produce proper frequency-dependent damping (see Section 2.2.2). (£) for the representation of long lines in lightning surge studies if no reflection comes back from the remote end during the duration tuax of the study. Example (f) 1s easily derived from Eq. (1.6a) if it 1s assumed that the initial conditions on the line are zero. In that case, hist ,.(t-1)=0 for t) Ce) (a) (e) (f) Among other things, self inductances are used to represent single-phase shunt reactors and neutral reactors in shunt compensa~ tion schemes (Fig. 2.4), part of discharge circuits in series-capacitor stations, equipment in HVDC converter stations, such as smoothing reactors, anode reactors, parts of filters on the ac and dc side, Inductive part of source impedances in Thevenin equivalent circuits for the “rest of the system” when positive and zero sequence para~ meters are identical (Fig. 2.5), Inductive part of single-phase nominal circuits in the single- phase representation of balanced (positive or negative sequence) operation or of zero sequence operation (Fig. 2.6), part of equivalent circuit for loads (Fig. 2.7), even though load modelling at higher frequencies is a very complicated topic [9], and loads are therefore, or for other reasons, often ignored, Fig. 2.5 ~ Thevenin equivalent circuit with Z,, os neg zero an ye generator trans~ Line as cascade connection of shunt former poninal tclreuits reactor Fig. 2.6 - Typical positive sequence network representation Fig. 2.7 - Load model for harmonics studies [9] (g) part of surge arrester models to simulate the dynamic characteristics of the arrester [10], (hn) parts of electronic circuits. 2-8 Choke coils used for power-line carrier communications are normally ignored in switching surge studies, but may have to be modelled in studies involving higher frequencies. Current transformers are usually ignored, un~ less the current transformer itself is part of the investigation (e.g., in studying the distortion of the secondary current through saturation effects). ‘The equation of a self inductance L between nodes k and m is solved accurately in the ac steady-state solution with Eq. (1.11). The only pre~ caution to observe is that ab, should not be extremely small, for the same reasons as explained in section 2.1.1 for the case of small resistance values. For the transient simulation, the exact differential equation at, ken % a ae (2.2) is replaced by the approximate central difference equation F———— TS z a ‘The same difference equation is obtained if the trapezoidal rule of integra~ 5 (2.3) tion is applied to the integral in it tygit? *tyg(t-ae +E fo [yy(u) - v,(w) Jee, (2.4) tar siving tyg(t) = Ayg(t—de) + FE fu Ct) ~ vaCe) + vj (eae) ~ vy(t~ae) J. (2.5) Eq. (2.3) and (2.5) can be rewritten into the desired branch equation at Syg(t) = Ze (y(t) ~ vale} + hist, (t-Ar), (2.6) with the “history term” hist,,(t-st) known from the solution at the preceding time step, ae hist, (t-ae) = dp a(ende) + oF fy, (tm de) - vy (e—at) je 7 Eq. (2.6) can conveniently be represented as an equivalent resistance R 2L equiv = =F, in parallel with a known current source histy,(t~At), as shown in Fig. 2.8. Once all the node voltages have been found at a particular time step at instant t, the history term of Eq. (2.7) must be updated for each hist, (t-Ae) Fig. 2.8 - Equivalent resistive circuit for transient solution of lumped inductance inductive branch for use in the next time step at t + At. To do this, the branch current mist first be found from Eq. (2.6), or alternatively, if both equations are combined, the recursive updating formula hist (t) = TE {u(t) - va(t)} + hist, (e-ae) (2.8) can be used. If branch current output is requested, then Eq. (2.6) is used. 2.24 Since the differential equation (2.2) is only solved approximately, it 4s important to have some understanding about the errors caused by the appli- cation of the trapezoidal rule of integration. As explained in Section 1-4 of Appendix I, the trapezoidal rule is numerically stable, and the solution does therefore not “run away” (see Fig. I.4 in Appendix I). Fortunately, there is also a physical interpretation of the error, because Eq. (2-5) resulting from the trapezoidal rule 1s identical with the exact solution of the short-circuited lossless line in the arrangenent of Fig. 2.9. This vas first pointed out to the writer by H. Maier, Technical University Stuttgart, Germany, in a personal communication in 1968, for the case of a shunt inductance. From a paper by P.B. Johns [12], it became obvious that this identity is valid for any connection of the inductance. To derive the parameters of such a “stub-line” representation, it is reasonable to start with the requirement that the distributed inductance L', multiplied by the 2-10 etub-line length 2, should be equal to the value of the lumped inductance : 2 Legere «<< a =f tttt - _— _ ay iy —e = ¢ Fig. 2.9 - lumped inductance rgplaced by short-circuited stub-line with 2 = 24 and t= ft With L'g known, the next parameter to be determined is the travel time 1 of the etub-line. Since = TDD, (2.10) the shorter the travel time, the smaller will be the value of the “parasitic” but unavoidable capacitance C't. The shortest possible travel time for a transient simulation with step size at is + Qa) with thts value, conditions at terminal 1 at £ ~ dt arrive at the shorted end at t= HE and got reflected back to terminal 1 at t. Therefore, the best possible stub-line representation has 2b be zeB wrk (2.12) Assume that the smoothing reactor on a de line has L = 0.SH, and that the step size is 100 js. Then Z = 10 000 9, and the unavoidable total capacitance C'£ becomes 5 nF, which appears to be negligible, at least if the reactor has a shunt capacitor of 1.2 uF connected to it anyhow, as in the case of the HVDC Pacific Intertie [11]. Now it remains to be shown that the exact solution for the lossless stub-line with parameters from Eq. (2.12) is identical with Eq. (2.5). As explained in Section 1, the expression (v + Zi) along a lossless line for a ficticious observer riding on the line with wave 2a speed remains constant, or going from 1 to 2 in Fig. 2.9, ae w(t - a) + 24g (e = at) = zig(e - , and travelling back again from 2 to 1, ae waig(e - AE) = vy(e) - 2, C0), which, combined, yields 1 AC) = Zvy(O + ( Z vyCeraey + Cea}, (2.13) which is indeed identical with Eq. (2.5). This identity explains the numeri- cal stability of the solution process: The chosen step size may be too large, and thereby create a fairly inaccurate stub-line with too much para~ sitic capacitance cory? on = Ge (2.14) * from Eq. (2.10), but since the wave equation 18 still solved accurately”), the solution will not run away. The "mathematical oscillations” sometimes seen on voltages across inductances, and further explained in Section 2.2.2, are undamped wave oscillations travelling back and forth between terminals 1 and 2 (Fig. 2-9). The identity of the trapezoidal rule solution with the exact stub-line solution makes it easy to assess the error as a function of frequency [13]- Assume that an inductance L is connected to a voltage source of angular frequency w, through some resistance R for damping purposes. The transient simulation of this case will eventually lead to the correct steady-state solution of the stub-line (or not drift away from the steady-state answer if the simulation starts from correct steady-state initial conditions). This steady-state solution at any angular frequency w is known from the exact equivalent n-circuit of Fig. 1-2. By short-circuiting terminal 5, the input impedance becomes * except for round-off errors caused by the finiteness of the word length in digital computers, which are normally negligible. There is no interpolation error, which occurs in the simulation of real transmission lines whenever + is not an integer multiple of At (see Section 4.2.2.2). 2-12 1 ZLaput ies 5 (2.15) eries * 7 “shunt or with Eq. (1-16), a tanwo $5) gape 7 BL (2.16) z Therefore, the ratio between the apparent inductance resulting from the stub-line representation or from the trapezoidal rule solution, and the exact inductance becomes be ‘rapezoidal z 5 (2.17) ‘The phase error is zero over the entire frequency range. Since power systems are basically operated as constant voltage networks, it makes sense in many cases to assume that the voltage V,(jw) across the inductance is more or less fixed, and that the current 1, (jo) follows from it. If we compare this current of the stub-line representation or trapezoidal rule solution with the current of the exact solution for the lumped inductance, then we obtain the frequency response of Fig- 2.10, where the ratio (the reciprocal of Eq. (2.17)) is shown as a function of the Nyquist frequency fyyquise ~ 260" a This frequency is the theoretically highest frequency of interest for a step size At, amounting to 2 samples/cycle. From a practical standpoint, at least 4 to 8 samples/cycle are needed to reproduce @ particular frequency even crudely. From Fig. 2.10 or from Eq. (217) it can be seen that the error in the current will be -5.2% at a crude sampling rate of 8 samples/cycle, or -0.8% at a more reasonable sampling rate of 20 samples/cycle. Furthermore, Fig. 2.10 also shows that the trapezoidal rule filters out the higher fre~ 2-13 Terapezoidal Texact 1.0 ! 0.0 0.5 1.0 /fyvauist ed 2 samples/cycle Fig. 2-10 - Amplitude ratio I through an trapezoidal! Texact Inductance as a function of frequency quency currents, since the curve goes dow as the frequency increases, as pointed out by R.W. Hamming [14]. Because of the error ia Eq. (2.17), there is a anall discrepancy between the initial conditions found with Eq. (1-11), and the response to power frequency in the tine step loop. For 60 liz, this error would be 0.012% with At=100 ps, or 1.2% with At*l ms. It is debatable whether Eq. (2.17) should be used for the steady-state solution, instead of Eq. (1-11), to match both solutions perfectly. This issue appears with other network elements as well If a perfect match is desired, then it may be best to have two options for steady-state solutions, one intended for inittalization (using Eq. (2.17) ta this exanple), and the other one intended for steady-state answers at one or nore frequencies (using Eq. (1-11) im this example). Very large values of L are acceptable as long as (oL)? or FE 48 not larger than the Largest floating point nuaber which the conputer can handle. To obtain flux | =f vat across a branch, a large inductance can be added in parallel and current output be requested. The need for this may arise {f 2 2-14 flux-current plot is required for a nonlinear inductance. With L = 1010 n, 1010 ~ times the current would be the flux. way as small resistances (see Section 2.1.1). 2.2.2 Damping of. a While the trapezoidal rule filters out high-frequency currents in induc- tances connected to voltage sources, it unfortunately also amplifies high~ frequency voltages across inductances in situations where currents are forced into them. In the first case, the trapezoidal rule works as an integrator, for which it performs well, whereas in the second case it works as a differ- entiator for which it performs badly. The problem shows up as “numerical oscillations” in cases where the derivative of the current changes abruptly, e-g+, when a current is interrupted in a circuit breaker (Fig. 2.11). The exact solution for v, is shown as a solid line, with a sudden jump to zero zest of network Fig. 2.11 - Voltage after current interruption 2-15 at the instant of current interruption, whereas the EMTP solution is shown as a dotted line. Since vce) = ZE (ace) - dceaey) - vp (tae), (2.19) and assuming that the voltage solution was correct prior to current interrup- tion, it follows that v(t) = -v,(t - At) in points 2, 3, 4,... as soon as the currents at t - At and t both become zero; therefore, the solution for v; will oscillate around zero with the amplitude of the pre-interruption value. ‘There are cases where the sudden jump would be an unacceptable answer anyhow, and would indicate improper modelling of the real system. An example would be the calculation of transient recovery voltages, since any circuit breaker would reignite if the voltage were to rise with an infinite rate of rise immediately after current interruption. For a transient recovery vol- tage calculation, the cure would be to include the proper stray capacitance from node 1 to ground (and possibly also from 1 to 2 and from 2 to ground). On the other hand, there are cases where the user is not interested in the details of the rapid voltage change, and would be happy to accept answers with a sudden jump. A typical example would be sudden voltage changes caused by transformer saturation with two-slope inductance models for the nonlinear- ity, as indicated in Fig. 2.12, It should be pointed out that these "numeri- Fig. 2.12 - Voltage jumps caused by transformer saturation 2-16 cal oscillations” always oscillate around the correct answer (around zero in Fig. 2.11) and plots produced with a smoothing option would produce the correct curves. Nonetheless, it would be nice to get rid of them, especially since they can cause numerical problems in other parts of the network, as has happened occasionally in turbine-generator models [26, p- 45]. The “textbook answer” would be re-initialization of variables at the instant of the jump. This would be fairly easy if the equations were written in state-variable form [dx/at]*[A][x]- With nodal equations as used in the EMTP, re-{nitialization was thought to be very tricky, until B. Kulicke showed how to do it [15]. His method is summarized in Appendix II. Whether re-initialization should be implemented is debatable, since the damping method described next seems to cure this problem, and also seems to have a physical basis as shown in Section 2.2.3. V. Brandwaja [16] and F. Alvarado [17] both describe a method for damp- ing these “numerical oscillations” with parallel damping resistances (Fig. 2.13). For a given current injection, the trapezoidal rule solution of the parallel circuit of Fig. 2.13 becomes 2L R - 2 vee) = gb {ace ~ t¢eraey} - RFE vee-aey. (2.20) Cr rs Ry the Fig. 2-13 - Parallel damping Téa current impulse is injected into this circuit (in a form which the EMTP can handle, e-g-, as a “hat, with 1 rising linearly to I,,, between 0 and 217 At, dropping linearly back to zero between At and 2At, and staying at zero thereafter), then, after the impulse has dropped back to zero, the first term in Eq. (2.20) will disappear, and we are left with the second term which causes the numerical oscillations, v(t) = -a + v(t-at), with a-2t_y (2.21) K+ pt ae being the reciprocal of the damping factor. This oscillating term will be danped if a <1; Lt 18 shom in Fig. 2.14 for R= 10 + Zor a= ty, and for 8, 72 2 or a= 5. The osciilation would disappear in one time step for * R,~ Hor a= 0 (critically damped case)". If R, 18 too large, then the poe damping effect 1s too sual. On the other hand, if R, 1s reduced until it approaches the value 2% (1deal value for damping), then too auch of an error 1s Introduced into the inductance representation. Fig. 215 shows the magnitude and phase error of the impedance for R, = 4 7¢ and R= 8 ZF, as well as the magnitude error from &q- (2-17) which already exists for the Inductance alone with the trapezoidal rule. It is interesting that che magnitude error with a parallel resistance is actually slightly smaller Fig. 2.14 - Oscillating term [17]. Reprinted by permission of F. Alvarado “\the critically damped trapezoidal rule with R, = 2L/at ts identical with the backward Euler method, as explained in Appendix 1.9. 2-18 = no phase error for R, 2L, . “ phase ame Rea error (°) o “ ‘ 2L I 8 ae 16 magnitude error (4) o “yquist Fig. 2.15 - Phase and magnitude error with parallel resistance [16] than the error which already exists for the inductance alone because of the trapezoidal rule. Therefore, the parallel resistance has no detrimental effect on the magnitude frequency response. It does introduce losses, however, as expressed by the phase error. As shown in the next section, these losses are often not far off from those which actually occur in equipment modelled with inductances. From a purely numerical standpoint, a 2-19 good compromise between reasonable damping (R, as low as possible) and acceptable phase error (R, as high as possible) leads to values of R, ‘series t 20 000 1 (2) ! ! | trapezoidal 5000 Yseries ™ 2.0, Hyyauise | ' Fig. 2.16 - Apparent series resistance and series inductance for the paratlel connection of Fig. 2.13, with e au region to the right of f/fy, Ste of ae practical interest because Ue S. SSepling rate would be too low to show these frequencies adequately 2-20 2L 2L 5.4 ESR, < 9.4 FF according to Brandwaja [16] (2.22) 20 | 2L or pt Te according to Alvarado [17], (2.23) with Brandwajn's lower limit determined by a specified acceptable phase error at power frequency. The errors introduced into the parallel connection of Fig. 2.13 through +x . series * series R,(5/(R, + $0 {8 show for the exact solution with X= ul, and for the trapezoidal rule solution with obj oezoigai fom Bas (2617). Whether the EMTP will be changed to include parallel resistances the trapezoidal rule are seen in Fig. 2.16, in which R automatically remains to be seen. It is interesting to note that the electronic analysis program SYSCAP of Rockwell International Corp., which Seems to use techniques very similar to the EMTP, has R, and R, of Fig. 2.18 butlt into the inductor model, with default values of R,=0-10 and B,=10!29 [22, p. 715]. The possibility of numerical oscillations is mentioned as well, in cases where the time constants of the inductor model of Fig. 2.18 are small compared with At (22, p. 773]. 2.2.3 Physi lel Rest There are many situations in which inductances should have parallel 1_Rei resistances for physical reasons. In some cases, the values of these resistances will be lover than those of Eq. (2.22) or (2.23), which will make the damping of the numerical noise even better. Typical applications of damping resistances are described next. These examples may not cover all applications, but should at least be representative. (a) Short-cireuit impedance of transformers The short-circuit impedance of transformers does not have a constant L/R-rati 3 Instead, the L/R-ratio decreases with an increase in frequency, as shown in Fig. 2.17 taken from [18]. If we use the curve for the 100 MVA transformer, and assume L = 1H (or mH, or p.u-) as well as At = 100 us, then a value of R, = 163 000 @ (or a, of prs) will produce the proper L/R-ratio at 1 kHz. This value ltes nicely in between the limits of 108 000 @ and 188 000 @ recommended in Eq. (2-22). 2-21 L E (as) 1 t | | poo ot : oo ‘ ° t | 3 : + : ' tot | oos sof 4923 5 10 0* > £ nz) Fig. 2.17 - L/R-ratio of the short-circuit impedance of typical transformers [18]. Reprinted by per- mission of CIGRE Fig. 2.18 - Equivalent circuit for the short-circuit impedance of a transformer A series resistance of R, = 9.4 9 (or Q, or peu.) can then be added to obtain the correct L/R-ratio at 50 Hz, which leads to the equivalent circuit of Fig. 2-18 for the short-circuit impedance of the transformer. 2-22 Wh Zigpue fom Eq. (2-16), thE Heegpezoigan/RFatt© of thts equivalent circuit is shown as a dotted Line in Fig. 2.19, which ts a reasonably good match for the experimental curve (solid line), and mich better than a constant L/R-ratio without R,*), It 1s interesting that a CIGRE Working Group on Interference Problems recommends the sane equivalent circuit of Fig. 2.18 for the analysis of haraonics [9], with ya . 13< ed < 30 (2.26) 8 0.002 0.0002 50 100 1000 "10 000 £(Hz) Fig. 2.19 - L/R-ratio of the short-circuit impedance of a 100 MVA transformer (dotted line from equivalent circuit of Fig. 2.18 with L being solved by trapezoidal rule, solid line from [18]). Reprinted by permission of CIGRE ")In the EMIP, the L/R-ratio without R, would actually increase with Frequency, since Loe oigay Of #4: (2-17) increases with frequency. and (b) (e) (4) 2-23 <0, (2.25) where Sy is the rated power and Vy the rated voltage of the transformer. +05 to 0-10 pu. at 50 Hz ([9] talks about ve ee cere hor eeetreuie 50 Hz), then Eq. (2.24) becomes with ul=(0.05 to 0.10)+——, N (40 841 to 81 6B1)+L ¥,) «ck (2.27) te replaced by the approximate central difference equation Ce) # 1 (ena) {v(t) ~ vCe) Joy, (eae = vg (t-at)} ee 2.28) aE which gives the desired branch equation 4 (e) = 25 fv (ey - vat} + nist, (e-ae) (2.29) a me f en p with the “history term” hist, (t-At) known from the solution at the preceding time step, hist, (t-at) [=] ob: be Requiv ~ 2¢ Fig. 2.22 - Equivalent resistive circuit for transient solution of lumped capacitance 2c bhet (EB) - ty Ce=Be) — FE (wy(erae) - vy(e-at)}- (2-30) Again, analogous to the inductance, identical results would be obtained from an integration of Eq. (2.27) with the trapezoidal rule. Eq. (2-29) can be at represented as an equivalent resistance 8,44;, "77 in parallel with a know current source hist, (t-At), as shown in Fig. 2.22. Once all the node 2-28 voltages have been found at a particular time step at instant t, the history term of Eq. (2.30) mist be updated for each capacitive branch for use in the next time step at tt\t. To do this, one must first find the current from Eq. (2-29). Alternatively, the recursive updating formula nto (2) = — Fe (ye) - vy(td} = nist, (eae) (2.31) can be used, which is the same as Eq. (2.8) for the inductance if followed by a sign reversal. Fig. 2.23 - Lumped capacitance replaced by stub-line with Zeat/2C and r=At/2 ys! Not surprisingly, the error analysis is analogous to that of the inductance. For a physical interpretation of the errors, the stub-line representation of Fig. 2.23 1s used, in which the lumped capacitance ts replaced by an open-ended lossless line. To obtain the parameters, it is reasonable to make the total distributed capacitance equal to the lumped capacitance, c= c. (2.32) With C'R known, the next parameter to be determined is travel time t. Eq. (2.10) shows that the shorter the travel time, the smaller will be the value of the "parasitic" but unavoidable inductance L't. For a step size At, the shortest possible travel time ts (2.33) utten Bq. (2.32) and (2.33) the surge tnpedance becones 2=88, 2-29 Without going through the details, let it simply be said that the exact solution for the stub-line of Fig. 2.23 is identical with the trapezoidal tule solution of Eq. (2-29) and (2.30). This identity will again be used to assess the error as a function of frequency. Assume that a capacitance C is connected to a source with angular frequency w, through some network with damping. The transient simulation will then settle down to the correct steady-state solution of the stub-line of Fig. 2.23, or not drift away from it {f the simulation was started from correct steady-state initial conditions. This steady-state solution is known from the exact equivalent n- circuit of Fig. 1.2, with terminal 5 being open-ended, 2 Ls Yeeries Yioput 7 series * 7 ‘shunt? ~ Hae tg eee 2.34) eries ~ Z ‘shunt or after sone manipulations with Eq. (1-16), tan(ap*) Yygput 7 S86 * (2.35) or This is analogous with Eq. (2-16) for the inductance, except that the analogous error now applies to the capacitance C rather than to the inductance L, or 1.0 Verapezoidal exact 0k 0.5 0.0 0.5 10 /fyauist . 8 4 2 samples/cycle Fig. 2.24 - Amplitude ratio Virapezoidal/Vexact Of a capacitance as a function of frequency 2-30 ae ¢, eancuft erapezotdar _ tay) ae + (2.36) mia Again, the phase error is zero over the entire frequency range. If we force a current I,(Ju) into the capacitance, then the voltage across the stub-line, compared with the exact solution, will have the frequency response of Fig. 2.24, which is identical with Fig. 2-10 if the current ratio is replaced by the voltage ratio. ‘The trapezoidal rule filters out the higher frequency voltages. Again, there 1s a small discrepancy between the initial conditions found with Eq. (1.12), and the response to power frequency from Eq. (2-35) in the time step loop (at 60 Hz, 0.012% error with At=100 ys, or 1.2% with At=l ms). Whether it should be eliminated has already been discussed in the second- paragraph of Section 2.2.1. Very small values of C are acceptable as long as (2p)° or $5 ta not larger than the largest floating point nunber vhich the computer can handle. —————— restetances (see Section 2.11), but they are unlikely to occur in practice. 2.3.2 While the trapezoidal rule filters out high-frequency voltages across capacitances for given current injections, it also auplifies high-frequency currents for given voltages across C. The nunerical oscillations discussed for the inductance in Section 2.2-2 would appear in capacitance currents 1f there 1s an abrupt change in dv,/at. For sone reason, numerical oscillations have seldom been a problem in capacitances, either because there are very few situations where they would appear, or simply because currents through Fig. 2.25 - Series damping resistance 2-31 capacitances are seldom included in the output. Analogous to the inductance, these numerical oscillations could be damped with series resistances R, (Fig- 2.25). Using Alvarado's arguments [17], the trapezoidal rule solution for a voltage impulse applied to the circuit of Fig. 2.25 would be ae £(¢) = pet {w(e) - veeraey} = FE * sce-aey. (2.37) +R, +R wt, wt After the voltage impulse v has dropped back to zero, we are left with the second term, which causes the numerical oscillations, A(t) = -a + A(t-At), At (2.38) - 8, witha = Zs wt, In analogy to Eq- (2.23), a reasonable value for the damping resistance would be e At R= 0.15 55 (2.39) 2.3.3 Physical Reasons for Series Resistance None are known to the writer at this time which would justify a series resistance as high as that of Eq. (2-39). G.W-A- Dummer [23] suggests the equivalent circuit of Fig. 2.26, and says that R, is dominant at very high frequencies, while R, te dominant at very low frequencies, but his comments Fig. 2.26 - Equivalent circuit for capacitor with losses 2-32 refer to capacttors used in electronics. ‘The typical textbook circuit has no series resistance, vhich would imply that the loss factor decreases inversely proportional with frequency. This contradicts the curve in Fig. 2.27 given by A. Roth for high-voltage capacitors [24]. Assuming C=1 uF and At=100 ys and using Ceraperoidar Umstead of C to duplicate the EMTP behaviour, value of RS =0.344 Q (ignoring R ‘> would more or less match tané at 50 Hz, while a value of cele .85+1077 § (ignoring Rg. ‘) would more or less match tan6 at 2 kHz, as shown in Fig. 2.27. Note that this value of R, is one order of magnitude tané (°/o9) 10 100 1000 —> £ (Hz) experimental ---- with R, and R P 's Fig. 2.27 - Loss factor [24]. Reprinted by permission of Springer-Verlag and A.W. Roth lower than the recommended damping resistance of Eq- (2.39). SYSCAP, an electronic analysis program with solution techniques similar to the ENTP, has R, and R, of Fig. 2.26 butte toto the capacitor model, with default values of Ro. _ prlol2 2 122, p 725]- Note that capacitors which may be subjected to short-circuits often have series resistors built in. Similarly, the overvoltage protection of series capacitors with spark gaps (Fig. 2.28) includes current-limiting R-L elements in the discharge circuit, with a typical “ringing” frequency of 400 Hz during discharge. 2 33 Fig. 2.28 - Spark gap protection of series capacitor 2.3.4 Example Nets ork with Capacitences Let us modify the fault current study of Section 2.2.4 for a case in which the transmission line is series 2.29). Let us further re compensated with capacitors (Fig. jeune that L, in Section 2.2.4 represented the net etance X,,,nul-1/uC at 60 Hz, to make both results directly comparable. With R, = 0.18 piu, X, © 1-0833 pus, uC = 2-695 peu. and at = 100 us, the fault current of Fig. 2.30 is obtained (data taken from [74], with connection from fault location to infinite bus left off). For comparison purposes, the done in Section 2.2.4, 46 shown as well; it differs appreciably from the more accurate solution with the circuit model of Fig. 2.29. This difference may fault current with the net reactance represented by L,, , 5 c closes Vax Sin (ot) { at to = + Fig. 2-29 - Single-phé -to-ground fault in a system with a series capacitor have some consequences for the accuracy of stability simulations, since net reactances are practically always used in stability studies. Fig, 2.32 compares the swing curves obtained with a net reactance and an L-C representation for a case similar to the IEEE benchmark model for sub- synchronous resonance studies (Fig. 2.31 differs from the figure shown in 1211). As can be seen, the differences are minor, possibly because there are many other reactances besides the line reactance in practical cases, which reduces the influence of the series capacitor. —t (ns) 2.0 Fig. 2.30 - Fault current in series-compensated network of Fig. 2.29 (line without symbols). For comparison, results fron Fig. 2.21 with met reactance are shown as well (1ine with ‘syabols) TEEE Subsynchronous Resonance Benchnark 4.09 | Scale 1000¢1) 4 mach 4) BETA 2 nacH | BETA 3.09. ee RL circuit RI-C circuit 1.00. 0.60 6.28 6.40 6.60 6.00 1.00 1.20 1.40" 1.68 Time scale: 18-0(8) 5. Fig. 2.31 - Swing curves with R-L and R-L-C representations 2-35 2.4 Series Connection of R, L, © Té lumped elements R, L, C frequently occur in pairs as series connections of R-L, R-C, or L-C, or as a series connection of all three elements R-L-C, then it becomes more efficient to treat the series connection as a single branch, thereby reducing the number of nodes and nodal equations. This has been tmplemented in the EMIP for the series connection of R-L-C (Fig. 2.32). For the steady-state solution, the branch equation is simply 1 Tea" RERGE STRY n?* k OA SIN Fig. 2-32 - Series connection of R, L, C To derive the branch equation for the transient simulation, add the three voltage drops across R, L, and C ‘6 - vy Rt yt Yer with the voltage drops expressed as a function of the current with Eq. (2.1), (2.6) and (2-29), QL , at 2 oe 7 v(t) ~ vyle) = OR + FE + FE) ty t) ~ AE hist, (e-we) — ye hhste(e-ae)- (2.40) After replacing the history terms hist, and hist, with the expressions of Eq. (2.7) and (2-30), this leads to the branch equation Syl) * ScrpenlMeCtVQCE] + MEE sores ESE» (2-414) L with eries ~ WE (2.41b) + 2-36 and the combined history term DLS, tog (t~SE“Cectog l(FE Re GE)MCt-at)+y, (tAt)=v, (tH AL)=2vg (EAE) } + (2.42) For updating this history term, the new current is first calculated from Eq. (2.41a), and the new capacitor voltage ve from volt) = vg(t-ae) + FE{i(e) + 1Ct-at)}. Eq. (2.42) is not the only way of expressing the combined history term, but it is the one being used in the EMTP. 2.5 Single-Phase Nominal n-Circuit This ts a special case of the M-phase nominal x-circuit discussed in Section 3.4. Earlier EMTP versions recognize the special case of M-l, and use scalar equations in place of matrix equations, whereas newer EMTP versions go through the matrix manipulations with M=l. Since single phase n-circuits are seldom used, it is reasonable to eliminate the special code for the scalar case. 31 - LINEAR, COUPLED LUMPED ELEMENTS Coupled lumped elements appear primarily in the M-phase wcircuit representation of transmission lines, in the representation of transformers as coupled impedances, and as source impedances in cases where positive and zero sequence parameters are not equal. 3.1 Coupled Resistances [RJ Coupled resistances, in the form of branch resistance matrices [R], appear primarily (a) as part of the series impedance matrix in M-phase nominal 1 ctreuite, (b) as long line representations in lightning surge studies if no reflections come back from the remote end during the duration t,,, of the study. The diagonal elements of [R] are the self resistances, and the off- diagonal elements are the mutual resistances. The off-diagonal terms in the series resistance matrix of an M-phase line are caused by the presence of the eareth as a potential current return path. The earth is not modelled as a conductor as such; instead, it is used as a reference point for measuring voltages. If it were explicitly modelled as a conductor, its equation for a three-phase line could have the form ay, ~ ae 7 “hala * “nts * 7ecTe * “he te Since the voltages are measured with respect to earth, V,"0, and, therefore, 2 2 2 EA EB EC ee ee een eee eT, Ta Ze 1B Zy Te' which, when inserted into the voltage drop equations for the phases A, B, C, produces av, Zh, 2 2h, 2h 2h, ay whe age BEBE, + age “ABBE + (oy “AB He = "hE EE aE Wor 32 and simtlar for B, C. This is the form used in M-phase reircuits, with earth being an implicit, rather than explicit, current conductor. Assuming purely Anductive coupling 2/,5X},, the terms Zi,Zfp/Zfp ete. will obviously contain real parte since the self impedance of the earth Zi, contains a real part. Whether the real part thus produced can strictly be treated as a resistance for all frequencies is open to debate, as explained in Section 4.1.24 The EMTP automatically converts a long line with distributed paraneters into a shunt resistance matrix [8] if (1) 1 ty, for all M modes”? of the M-phase Line, (2) zero initial conditions. This representation 1s simply an N-phase generalization of the single-phase case discussed in Section 2.1. For the high-frequency lossless line model, which de often used in lightning surge studies and described in more detail 4m Section 4.1.5.2, this shunt resistance matrix has the elenents = 60 mat, Ry = 60 in at GD R : 1 ik 1 with hy-average height above ground, r,=conductor radius, Dj,=distance from conductor i to image of conductor k, d,,sdirect distance between conductors 1 and k, ‘These are the well-known self and mutual surge impedances of an M~ phase line [8]. The equations for coupled resistances Myg(t)] = (RIT? {Iv (0)] - tv 6e) 1} (3.2) are solved accurately by the EMTP, as long as [R] is non-singular and not extremely ill-conditioned. In all cases known so far, [R] is symmetric, and the EMTP has therefore been written in such a way that it only accepts symmetric matrices [R]. ‘The EMTP does not have an input option for coupled resistances by themselves; instead, they mist be specified as part of the M-phase nominal weireuit of Section 3.4, with L and C left zero. For long lines with +> tygyand zero initial conditions, the EMTP converts the distributed- YModes are explained in Section 4.1.5. 33 parameter model internally to the form of Eq. (3.2). Since [R] is symmetric, the EMTP stores and processes the elements of these and all other coupled- branch matrices as one-dimensional arrays in and above the diagonal (e.g. Ry, stored in X(1), Ry, im X(2), Ryy in X(3), Ry, in X(4), ete). Bold 34 As already mentioned, [R] must be non-singular if a resistance matrix is read in, If its inverse [R]~! 16 read in, then this requirement can be dropped, since [RJ-} is allowed to be singular without causing any problems. Also, the resistances shouldn't be so small that [R]~! becomes so large that 4t “swamps out” the effect of other connected elements, as mentioned in Section 2.1.1, On the other hand, very small values of [R]~! are acceptable (see “very large resistances” in Section 2.1.1). Since coupled branches have not been discussed in the introduction to the solution methods, their inclusion into the system of nodal equations shall briefly be explained. Assume that three branches ka-ma, kb-ab, ke-mc are coupled (Fig. 3.1). In forming the nodal equation for node ka, the current 1,4 gq 18 needed, ct im Fig. 3.1 - Three coupled resistances branch branch branch, Seana "San ka Yaa) * Cab kbub) * Sac (ke%ne)* 4 with GPraPCh being elements of the branch conductance matrix [R]~!. This means that in the formation of the nodal equation for node ka, GF" enters into clement Gp, 4 of the nodal conductance matrix in Eq. (1.88), acbranch Fanch branch = Ant Gana? Say AEO Ge yyy “Gop dnto Gays etc. If this is done systematically, the matrix [R]~! will be added to two diagonal blocks, and subtracted from two off-diagonal blocks of the nodal conductance matrix [G], as indicated in Fig. 3.2. Unfortunately, rows ka ke oma ne WV I I tory te ka—| ot oH : eee -tRy2 arty} Fig. 3.2 - Contributions of three coupled branches to the nodal conductance matrix and columns ka, kb, ke and ma, mb, mc do not follow each other that neatly, and the entries in [G] will therefore be all over the place, but this is simply a programming task, It is worth pointing out that the entry of coupled branches into the nodal conductance matrix can always be explained with an equivalent network of uncoupled elements. For three coupled resistances, the equivalent network with uncoupled elements would contain 15 uncoupled resistances (see Fig. 7 in Chapter II of [26]). Such equivalent networks with uncoupled elements are useful for assessing the sparsity of a matrix, but they can be misleading by seemingly indicating galvanic 35 connections where none exist. For example, the steady-state branch equations for two-winding transformers, which are well known from power flow and short~ circuit analysis, Txa,ma| Yana - . (3.3) Fepymb} [FY YY [Meo Ya| simply imply the connection of Fig. 3.3(a), and nothing more. The equivalent network with uncoupled elements is shown in Fig. 3.3(b), which produces the well-known transformer model of Fig. 3.3(c) if nodes ma and mb are grounded. ka tY kb ty Ne ond joo ma tY mb (a) Coupled elements (b) Uncoupled elements (ce) Uncoupled elements with nodes ma and mb grounded a & Fig. 3.3 - Two-winding transformer as two coupled branches 3.1.3 Example for Coupled Resistances Assune that a lightning stroke, represented by a current source i(t), hits phase A of a three-phase line (Fig. 3.4). Let us then find the voltage +6 itt) Fig. 3.4 - Lightning stroke to phase A of a three-phase line build-up in all 3 phases over a time span during which reflections have not yet come back from the remote ends of the line, using the high-frequency lossless line model of Eq. (3-1). Assume a flat tower configuration typical of 220 kV lines, with an average height above ground=12.5 m, spacing between conductors=7.58 m, and conductor radius=14.29 mm. Then from Eq. (3-1), $48.02 74.26 39.41 (R] = |] 74.24 448.02 74.24) Qe 39.41 74.24 448.02, The left as well as the right part of the line is then represented by [R] connected from A,B,C to ground, and the voltages become va(t) = 224,01 4(8), vg(t) = 37-12 A(t), volt) = 19.71 4(t), or 16.6% of v, appears in phase B, and 8.8% in phase C. An interesting variation of this case is the calculation of the effect which this lightning stroke has on the equipment in a substation. Assume that the travel time t between the stroke location and the substation is such that no reflection comes back from the stroke location during the time t... of the study, with the time count starting when the waves hit the substation (Fig. 3.5). In such cases, the waves coming into the substation can be represented as a aT three-phase voltage source with amplitudes equal to twice the value of the voltages at the stroke location, behind the resistance matrix [R]- This, in turn, can be converted to a current source in parallel with a shunt resist- A(t) Usouccal = 218) | |) 3 simply becomes equal to the lightning current at the stroke location [i(t), from the stroke location. (a) Network configuration AQ), ie (b) Equivalent network for line and lightning stroke Fig. 3.5 - Waves coming into substation 38 3.2 Coupled Inductances [1] coupled inductances, in the form of branch inductance matrices, are used to represent magnetically coupled circuits, such as (a) inductive part of transformers, (b) inductive part of source impedances in three-phase Thevenin equivalent circuits for the "rest of the system” when positive and zero sequence parameters differ, (ce) inductive part of M-phase nominal tcircuits The diagonal elements of [L] are the self inductances, and the off— diagonal elements are the mutual inductances. In all cases known so far, [1] is symmetric, and the EMTP only accepts symmetric matrices, with the storage ‘scheme described in the last paragraph before Section 3.1.1. ‘The source impedances mentioned under (b) above are often specified as positive and zero sequence parameters 2,45 Zzerq which can be converted to ‘pos’ “zero self and mutual impedances 1 1 24° 52 p06™ zero)? *a ~ 3(zer0™*po8) (4) of the coupled impedance matrix Z\° (3.5) Of course, these self and mutual impedances can in turn be converted back to sequence parameters, Zoos = Zen2m? Znero 7 Zet@m * (3.6) ‘pos a? “sero” “oo For a generalization of thie data conversion to any number of phases M, see Eq. (4.60) in Section 4.1.3.2+ +9 ‘The equations for coupled inductances between a set of nodes ka, kb, -.+ and a set of nodes ma, mb, ... (Fig. 3.6) are solved accurately in the ac steady-state solution with i =i eq) * FH GT {do Lvg) (G.7) The only precaution to observe is that [L]-} should not be extremely large, for reasons explained in Section 2.1.1. Ye hate ke arin me ka oo ma L----4 ita) Fig. 3.6 Four coupled inductances For the transient simulation, Eq. (2.6) and (2.7) for the scalar case are simply generalized for the matrix case, which produces the desired branch equations (yg) = ELIT ly (ed) — Uv (ed] } + Inst gta], G8) with the history term [hist,,(t-6t)] known from the solution at the preceding time step, Inter, (tard) = (yg (e-aed] +E te) {ly (eae) - Iv g(e-aed}}. 3.99 Just as in the uncoupled case, Eq. (3.8) can be represented as an equivalent resistance matrix [Rogujy/"(2/st){L], in parallel with a vector Intst,.,(t-4t)] of known current sources. The matrix [R,o41¥! the nodal conductance matrix of the transient solution in the same way as J7} enters into described in Section 3.1.2 (for the steady-state solution, simply replace IR, ‘equiv inductance enters only into two components k and m of the right-hand side in 7) by (1/ju){L]7}). While the current source hist,, of an uncoupled 3-10 Eq. (1,8), the vector [hist,,] must now be subtracted from components ka, kb, ke, ... , and added to components ma, mb, mc, +. Once all the node voltages have been found at a particular time step at instant t, the history term of Eq. (3.9) must be updated for each group of coupled inductances. This could be done recursively with the matrix equivalent of the scalar equation (2.8). The EMTP does not have an input option for coupled inductances alone; instead, they must be specified as part of the M-phase nominal rcircuit of Section 3.4, where the updating formulas used by the EMTP are discussed in more detail. There are situations where [L] may not exist, but where [L]~! can be specified as a singular matrix. Such an example is the transformer model of Eq. (3.3). If resistances are ignored, Eq. (3.3) can be used for transient studies with ay y -+F [lpg eaee ee (3.10) where Y = 1/(JX), with X being the short-circuit input reactance of the transformer measured from winding ka~ma, It is therefore advisable to have input options for [L]~} as well as for [L], as further discussed in Section 3.4.2. 3.2.1 Error Analysis The errors are the same as for the uncoupled inductance, that is, the ratio tan(w $)/(w$*) of Eq. (2-17) applies to every element in the matrix [1], or its reciprocal to every element in [L]~!. The stub-line representation of Fig. 2.9 becomes an M-phase stub-line, if M is the size of the matrix [L]. There 1s no need to use modal analysis for this stub-line because all travel times are equal, as mentioned in Section 4.1.5.2. In that case, the single-phase line equations can be generalized to M-phase line equations by simply replacing scalars with matrix quantities. Eq. (2.9), (2,12) and (2.14) therefore become atL'} = [1], (2) 2 . % (ul, ana s{c"] 31 3.2.2 Damping _of Again, the explanations of Section 2.2.2 for the uncoupled inductance are easily generalized to the matrix case if all elements of [L] are to have the same ratio R,/L. Since [L]~! 18 used in Eq. (3.8), it 1 preferable to express the parallel resistances in the form of a conductance matrix, e-g., with Alvarado's recipe of Eq. (2.23), Gen) [6,} = 0.15 5 (117? If (L]~! is singular, Is) would be singular as well, but the singularity would not cause any problems. If the coupled inductances go from nodes ka, kb, ... to nodes ma, mb, + (Pig. 3.6), then {e) would be connected in the same way from nodes ka, kb, +». to nodes ma, mb, The reasons are the same as those listed in Section 2.2.3 in those situations in which the single-phase case can be generalized to the M-phase 3.2.4 Example for Network with Coupled Inductances Let us go back to the example of the single~phase-to-ground fault described in Section 2.2.4, but treat it as a three-phase Thevenin equivalent circuit now, with coupled resistances and inductances (Fig. 3.7). Assume that Z,,, 7 0602 + 40-404 peus and 2,554 ~ 015 + 32-329 prus, or with Eq. (3.4) (2, = 0018 + 40.712 pous, Z, = 0416 + 40-308 peu. There are three voltage sources closes at t=0 Fig. 3.7 - Single-phase-to-ground fault with three— phase Thevenin equivalent circuit 312 7 Ya-source ~ Yaax®in(wt, Yp-source ~ YaaxSimut-120°), Veugource ~ YmaxSiM(wttl20°)- With the same values of R, and L, as in section 2.2.4, the fault current will be identical with the curves of Fig. 2.21(b) and (c). In addition, we can now obtain the overvoltages in the unfaulted phases B and C, which are shown in Fig. 3.8. o NIRS Www Fig. 3.8 - Overvoltages in unfaulted phases B and C The steady-state solution can of course be easily obtained from the phasor equations y Va-SOURCE 2, 2 2 an ‘ Yp-SoURCE eee ele (B.12a) Z, oO Vo-souRcE 2 z, ‘8 The first row produces, with V,=0, 1, = LArsouRce 120) A Zs +13 ‘and the second and third rows produce the voltage changes in the unfaulted phases z, iy = 86 =~ 2 Ya sounce" (3.120) If these voltage changes are shown in a phasor diagram (Fig. 3.9), then it becomes obvious why the overvoltages in pha: B and C are unequal, unle: the ratio Z,/Z, 16 a real (rather than complex) number. In the latter c the dotted changes become vertical in B and C in Fig. 3.9, and the over- voltages becone equal. —— pre-fault values ssees changes caused by fault o> fault values Fig. 3.9 - Phasor diagram of voltage changes caused by single-phase-to-ground fault 3.3 Coupled Capacitances [C] Coupled capacitances, in the form of branch capacitance matrices, appear 2 the shunt elements of M-phase nominal x-circuits (Fig. 3.10). One could argue that the capacitances are not really coupled, since they appear as 6 uncoupled capacitances in Fig. 3-10. However, the same argument can be made for coupled resistances and inductances, as explained in Fig. 3.3 of Section 3.1.2, and the fact remains that the shunt capacitances of M-phase lines appear as matrix quantities in the derivation of the equations. Fig. 3.10 ~ Three-phase nominal rcircuit Since the only Inova application of coupled capacitances 4 as shunt elesents of tephase nominal reizculte, the EATP accepts then only in that form, that 4a, en equal branch capacitance autrices 2 {C) at each end of the m-circuit, from nodes ka, kb, -.. to ground, and from nodes ma, mb, ++. to ground. In all cases, [C] is symmetric, and this symmetry is exploited with che storage schene deveribed in the 1 t paragraph before section 3.1.1. The Atagonal elenent C.,,, of [C] is the sum of all capacitances between phase 2 and the other phases b, c, ++. as well as between phase a and ground, whereas i: the off-diagonal element C,,,, 18 the negative value.” of the capacitance between phases a and by Sometimes, shunt capacitances of three-phase lines are specified as positive and zero sequence parameters C_,., C,..5 which can be converted to pos’ zero the diagonal and off-diagonal elements Fl sh . CF l,06 * Cero)? Sa F zero ~ “pos? (3.13) of the coupled capacitance matrix *)it might be worthwhile to have the EMTP check for the negative sign, and automatically make it negative, with an appropriate warning message, in cases where the negative eign was forgotten. The writer is not avare of any situation in which the off-diagonal element would not be negative. oo G [cl = Je, ce} (3.14) c, CG, C, ©, must be negative because the off-diagonal element is the negative value of the coupling capacitance between two phases, therefore, C.4,,6 C,g+ For @ generalization of this data conversion to any number of phases M, see Eq. (4,61) in Section 4.1.3.2. The steady-state equations for coupled capacitances in the shunt connection of Fig. 3.10, and with the factor 1/2, are os wi ol 7 ZIWICHYI, [yo] =F Jw (ollv,) (3.15) with subscripts “ko” and “mo” indicating that the currents flow from nodes ka, kb, «.. to ground ("0"), and from nodes ma, mb, ... to ground. Eq. (3.15) 18 solved accurately in the steady-state solution. The only pre~ caution to observe is that w[C] should not be extremely large, for reasons explained in section 2.1.1, but this is very unlikely to occur in practice anyhow. For the transient simulation, Eq. (2-29) and (2.30) are again general- ized for the matrix case, which produces the desired branch equations (taking care of the factor 1/2!), L4y.g(t)] = GE lolly, (ed) + Unter, (eae), G.16) with the history term (hist, (t-at)] known from the solution at the preceding time atep, 1 htet,,(e-ae)] = - sb fel lv, (emse)] = fay (omar) ] (3.17) The equations for the shunt capacitance 1/2 [C] at the other end (nodes ma, mb, ..+) are the same if subscript k is replaced by m. As in the uncoupled case, Eq. (3.16) can be represented as an equivalent resistance matrix ae(cl-1, in parallel with a vector [histy,(t~ét)] of known current sources. The matrix zt [C] enters into the nodal conductance matrix of the transient solution only in the diagonal block of rows and columns ka, kb, ... and in the diagonal block of rows and columns ma, mb, ... (Fig. 3.2), because of 3416 the shunt connection, while the vector [hist,,] must be subtracted from components ka, kb, ... (analogous for {hist .])« Once all the node voltages have been found at a particular time step at instant t, the history term of Eq. (3-17) 4s updated recursively, Uhisty.(t)} = - ge (C][y(t)] - (hist, (eaty) G18) and analogous for [hist,.]- Recursive updating ts efficient here, in contrast to coupled inductances, because the branches consist only of capacitances here, unless currents are to be computed as well. In the latter case, [1,,(t)] is first found from Eq. (3.16), and then inserted into Eq. (3.17) to obtain the updated history term, with both formulas using the same 1 matrix GF [C]- The errors are the same as for the uncoupled capacitance, that is, the ratio tan(w 4£)/(w 4) of Eq. (2.35) applies to every elenent in the matrix 1/2 [C]. The stub-line representation of Fig. 2.23 becomes an M-phase stub- lime, with the second set of nodes being ground in this case. There is no need to use modal analysis, as explained in Section 3.2.1. 3.3.2 Damping o: te Series _B 8 Again, the explanations of Section 2.3.2 for the uncoupled capacitance are easily generalized to the matrix case if all elements of 1/2 [C] are to have the same time constant RC. Eq. (2.39) would then becone [R,] = 0615 a [oy (3.19) (factor 1/2 of Eq. (2.39) disappeared because the equations have been written for 1/2 [C] bere). As mentioned in Section 2.3.2, mmerical oscillations in capacitive currents have seldom been a problen. 3-17 3.3.3) Bh pl 28 None are known to the writer at this time. The discussions of Section 2.3.3 do not apply to shunt capacitances of overhead lines, but they may be relevant to the capacitances of underground or submarine cables. 3.3.4 Example for Network ith Coupled Capacitances Assume that a power plant with a nuuber of generator transformer units in parallel is connected into the 230 kV switchyard through a number of parallel underground cables. The circuit breakers at the end of the cables are open, when a single-phase-to-ground fault occurs on the power plant side of the breakers (Fig. 3.11). eo read eo 4 Pig. 2.11 ~ cable efroute with slagle-phase-to-ground fault. Fault occurs in fhsee Athen source voltage in A iovat its peaks Generator trans foracrs represented 9 three-phase voltage, sources of 230 KV Gans, Sine tovtine) behind’ coupled’ reactances with Kove = 81, Les neon (retecred to 250 RY side). Cables veprESénted as THEEL prase oosinal wetreule with Zpoy = Zere” OCOLEIED, poy = = 897.6 US =la ls le ae > Sexour the data recenbles the situation at Grand Coulee before the Third Powerhouse was butte, except that Z,,., = 24, for the cables 1s an unrealistic zero * “pos ssnaption. Also note that the shone capecttances of the nominal w-circutt ees taste cot asee fences at teeing atarears et high voltage cables where each phase ts electrostatically shielded. Nonetheless, thts cable efrcult wae chosen because it illustrates the effects of shunt capacitances better than a case with overhead Lines where C, # 0. 318 v] w 200 (a) Overvoltages =1 | 40 ° 3 6 9 —— t (ms) 40 (b) Fault current for Rpgyyp " 1 @ (negative value shown) 3-19 —> t (ms) (ce) Fault current for Rpayyp = 0 (different scale than in (b), but value again negative) Fig. 3.12 - Overvoltages and fault current for a single-phase-to-ground fault in the cable circuit of Fig. 3.11 (tt = 10)s; small step size chosen to allow comparisons with distributed parameter model of cable with Zuoge 7 462%, t= 108) Fig. 3.12(a) shows the voltages in the two unfaulted phases at the fault location, with oscillations superimposed on the 60 Hz so typical of cable cireuits. Fig. 3.12(b) shows the fault current; the high-frequency oscillations at the beginning are caused by discharging the shunt capacitance through the fault resistance of 19, With zero fault resistance, this Gischarge would theoretically consist of an infinite current spike at t = 0, which leads to the undamped numerical oscillations across the correct 60 Hz ~ values discussed in Section 2.3.2 (Fig. 3.12(c)). These numerical oscillations would not appear if the cables were modelled as lines with distributed parameters; instead, physically based travelling wave oscillations would appear which would still look similar to those of Fig. 3.12(b). 3-20 3.4 M-Phase Nominal mCircuit Series connections of coupled resistances and coupled inductances first appeared as part of M-phase nominal tcircuits (Fig. 3.10) when the EMTP was developed. It was therefore decided to handle such series connections as part of an M-phase nominal rcircuit input option. By allowing the shunt capacitance 1/2 [C] to be zero, this circuit input option can then be used for series connections of [R] and [L] as well. The equations for the shunt capacitance matrices 1/2 [C] at both ends are solved as discussed in section 3.3. [C] = 0 is not recognized by the EMTP as a special case; instead, the calculations are done as if [C] were What remains to be shown is the series connection of [R] and [L] as one single set of M coupled branches. The derivation of the coupled branch equations is similar to that of the scalar case discussed in Section 2.4, if scalar quantities are replaced by matrices. When the series (R] - [L] connection was first implemented in the EMTP, it was not recognized that [L] may not always exist. With the appearance of singular [L]"} matrices, e.g-, in the transformer model of Eq. (3-3), an alternative formulation was developed. Both formulations have been implemented, as discussed in the next two sections. 3.4.1 Series Connection of [R] and [L] For the steady-state solution, the branch equations are - jo ]}-} - . 7 Tyg 7 (OR + Soh} Civ) - yd) 3.20) They are solved accurately. For the transient simulation, the branch equations are derived by adding the voltage drops across [R] and [L]. From Eq. (3.2) and (3.8), [yg (0] = [6 1 {ly,C0)] - (v,Ce]) + [hise, (ear)} (3.218) series: series’ with IR, and the history term series! ~ (I+ 3 [EIs 804 [6yer405) ~ Byerteg!? G22 3-21, (hist, (e-ae)] = (6, feertos (M(t BEd I-Lvg (ted HG [LJ series IRD [Cee » (3-22) Direct updating of the history tera with Eq. (3.22) involves three matrix nultiplications because [1,,] must first be found from Eq. (3.214). Unless currents must be computed anyhow, as part of the output quantities, updating with the following recursive formula is more efficient, [hist cepen(t)] = (HM Iv CE) ]-[vg(t) #18 gerteg] [nse “bey series series series‘ [hist, (eae) ] (3-234) series since it involves only two matrix multiplications. Matrix [H] is IH) = 2(16 - IG, TERIIG, De (3.23b) ‘series! ‘series ‘series’ All matrices [Reories!* (series exploited by the EMIP with the storage scheme discussed in the last paragraph ] and [H] are still symmetric, which is before Section 3.1.1. Symmetry is not automatically assured. For instance, the alternative updating formula ()] = CFI] - [hist thise cenaey), series! series! which, in combination with Eq. (3-2la), would be preferable in situations where current output is requested, has an unsymmetric matrix [F], ee tte ee ie seetiaaee All equations in this section can handle the special case of either (R] = 0 or [L] = 0 as long as [R, ] of Eq. (3-21b) can be inverted. ‘series {R]_and_[L]-) Singular matrices [{L]~! appear in transformer representations if 364.2 exciting currents are ignored. By itself, [L]"! ts easily handled with Eq. (3.8) and (3.9). In series connections with [R], however, the equations of the preceding section cannot be used directly because [L] does not exist. For the steady-state solution, the matrix [R] + jo [L] is rewritten as (R] + Jo [L] = (Sol) {T5oLI"? fk] + (UD, with [U] being the identity matrix, which upon inversion, produces the inverse required in Eq. (3-20), ({R] + go [L]}7) = (0) + [gon]! pRy? [jon]. (3.24) 322 Eq. (3.24) produces a symmetric matrix, even though the matrix [U]+{juL]~'{R] needed as an intermediate step is unsymmetric. The symmetry of the result from Eq. (3-24) can be shown by rewriting the matrix [R] + ju[L] as CR] + jw Ch) = (yok) gchIT URI aIWE + EobI“1) (9k), from which the inverse is obtained as (ER) + jw (L} 7} = (jab]7? (05 ob]- RNG ub)-? + C§ab~ 2}? (gubl“. (3,25) Each of the three factors of the product is a symmetric matrix, which is obvious for the two outer factors and which can easily be proved for the inner factor by showing that its transpose is equal to the original. With all three factors being symmetric, the triple product [A][B][A] is symmetric, too. The EMTP uses Eq. (3.24) rather than (3.25), because the latter would fail if [R] inversion, followed by matrix multiplication with an imaginary matrix, © and {LJ~} singular. ‘The EMTP does not use complex matrix however. Instead, Eq. (3.24) is reformulated as the solution of a system of N linear equations with N right-hand sides, {atu) + Cut) (R) tv) = Can]? (3.26) where the inverse [Y] is now directly obtained as the N solution vectors. To avoid complex matrix coefficients, Eq. (3.26) is further rewritten as 2N real equations, - 1 tal Rd.) = x) = bot (3.27) fy} + Coll (RLY, ]} = 0. (3.28) By replacing [¥,] in Eq. (3.27) with the expression from Eq. (3.28), the imaginary part of [¥] is found by solving the N real equations (oat) + tu pty) = = Cal (3.29) and the real part is then calculated from Eq. (3.28). For the transient simulation, [R, J of Eq. (3.21b) is rewritten as ‘series my +) = 2 uel + Wi}, which, upon inversion, produces the matrix [G, ] required in Eq. (3.21a), series 3423 (ceries P= + ey (3.30) z 2 Again, the matrix [U] +$© [LJ-1R] needed as an intermediate step is unsymmetric, while the final result [6,405 proved with Eq. (3.25) by simply replacing ju by 4. As in the steady-state ] becomes symmetric. Symmetry is case, the inverse of Eq. (3.30) is found by solving N linear equations {to + te} te. yeg! (3.31) series To initialize the history term [hist,..).,], Eq- (3+21a) can be used directly. To update it, neither Eq. (3.22) nor Eq. (3-23a) can be used because [L] and [R, ‘series! do not exist. Instead, Eq. (3.23a) is rewritten as Thistyorpeg(t)] = (Hl Ulvy(t)] ~ Uvg(t)1) + thse. og (emee)] + (3.32) Ic J [-2R] [hist gg(trae) de series By storing the symmetric matrices [H], [6,..,,,] and -2[R], the updating with Eq. (3.32) can be done with three matrix multiplications, starting with the product -2[R] [hist (t-at)], An alternative updating formula, which J and -2[R], and series requires the storage of only two symmetric matrices [G. 2405 produces the currents [1,,] a8 a by-product, is Intse, (e)] = (6, T flv Ce] ~ fC] + (-2R1 COT} + UC) series ‘series (3.33) if the current is first found from Eq. (3.2la), followed by the Semeebeeeer oes -2[R] aye, etc. Eq. (3.33) is derived from Eq. (3.22) by rewriting 7 [L] - [8] 25 [Rygrieg) ~2(RI- All equations in this section have symmetric matrices, and can handle the special case of either [R] = 0 or [L]~} = 0 as long as [U] + $= (L]~1(R] in Eq. (3.30) can be inverted. Note, however, that [L]~1=0 implies infinite inductances, that is, the M coupled branches are really M open switches. 4-1 4. OVERHEAD TRANSMISSION LINES 4.1 Line Parameters The parameters R', L', and C’ of overhead transmission lines are evenly distributed along the line”, and can, in general, not be treated as lumped elements. Some of them are also functions of frequency; therefore, the term “Line constants” is avoided in favour of “Line parameters”. For short- circuit and power flow studies, only positive and zero sequence parameters at power frequency are needed, which are readily available from tables in For the line models typically needed in EMIP studies, however, these simple formulas are handbooks, or can easily be calculated from simple formula: not adequate enough. Usually, the line parameters must therefore be computed, with either one of the two supporting routines LINE CONSTANTS or CABLE CONSTANTS. These supporting routines produce detailed line parameters for the following types of applications: (a) Steady-state problems at power frequency with complicated coupling effects. An example is the calculation of induced voltages and currents in a de-energized three-phase line which runs parallel with an energized three-phase line. Both lines would be represented as six coupled phases in this case. (b) Steady-state problems at higher frequencies. Examples are the analysis of harmonics, or the analysis of power line carrier communication, on untransposed lines. (c) Transients problems. ‘Typical examples are switching and lightning surge studies. Line parameters could be measured after the line has been built; this is not easy, however, and has been done only occasionally. Also, lines must *)the “prime” in R', L' and C' is used to indicate distributed parameters in O/km, H/km and F/km. 4-2 often be analyzed in the design stage, and calculations are the only means available for obtaining line parmeters in that c; ‘The following explanations describe primarily the theory used in the supporting routines LINE CONSTANTS and CABLE CONSTANTS, though other methods are occasionally mentioned, especially if it appears that they might be used in EMTP studies some day. The supporting routine LINE CONSTANTS is heavily based on the work done by M.H. Hesse [27], though some extensions to it were added. 4elel The solution method is easier to understand for a specific example. Therefore, a double-circuit three-phase line with twin bundle conductors and one ground wire will be used for the explanations (Fig. 4.1). There are 13 2 ase! (22 se} (2'] of (2) Le ite or or & E symmetrical B Sorponents & fora dephase cireuits 13 conductors Fig. 4.1 - Line parameters 4-3 conductors in this configuration. They will be called individual conductors”, to distinguish them from the 6 equivalent phase conductors which are obtained after pairs have been bundled into phase conductors and after the ground wire has been eliminated. 4.1.1.1 Series Impedance Matrix It 4s customary to describe the voltage drop along a transmission line in the form of partial differential equations, e.g-, for a single-phase line as ~ Meese Bb x Ri +L! a (4.1) ‘The parameters R' and L' of overhead lines are not constant, however, but functions of frequency. In that case it is improper to use Eq. (4.1); instead, the voltage drops must be expressed in the form of phasor equations for ac steady state conditions at a specific frequency. For the case of Fig. 4a, avy “& My By re Mayas |] avy ez Setter ee aes 19¥ ||| - - (4.2a) av 13 1 . . ax 231 713,207 213,13 ]L%3 with V, = voltage phasor, measured from conductor i to ground, Iy= current phasor in conductor 4, or in general -)- im (4.20) *)in the output of the,supporting routine LINE CONSTANTS, they are called “physical conductors". Anh with [V] = vector of phasor voltages (measured from conductor to ground), and [1] = vector of phasor currents in the conductors. Implied in Eq. (4.2) is the existence of ground as a return path, to which all voltages are referenced. The matrix (2'] = [R'(u)] + ju [L'(w)] is called the series impedance matrix; it is complex and symmetric. The Rig, + jwL',, is the series self impedance per unit length of the loop formed by conductor i and ground return. The off-diagonal element 2", diagonal element Z' ie = 2'gy 7 Bly + jul'yy is the series mutual impedance per unit length between conductors i and k, and determines the longitudinally induced voltage in conductor k if a current flows in conductor i, or vice versa. The resistive terms in the mutual coupling are introduced by the presence of ground, as briefly explained in Section 3.1. Formulas for calculating Z';, and Z',, were developed by Carson and Pollaczek in the 1920's for telephone circuits [28,29]. These formulas can also be used for power lines. Both seem to give identical results for overhead lines, but Pollaczek's formula is more general inasmuch as it can also be used for buried (underground) conductors or pipes. Carson's formula is easier to program than Pollaczek's and is therefore used in both supporting routines LINE CONSTANTS and CABLE CONSTANTS, except that the latter includes an extension of Carson's formula for the case of multilayer stratified earth [30] as well. Carson's, Pollaczek's and other earth return formulas are compared in [31]. Two recent new approaches to the calculation of earth-return impedances are the subconductor method of Hartenstein, Koglin and Rees [32], and the complex depth approach of Gary, Deri, Tevan, Semlyen and Castanheira [33,34]. Hartenstein, Koglin and Rees treat the ground as a system of conducting layers 1,2,3,...,n with uniform current distribution in each layer (Fig. 4.2(a)). Their results come close to those obtained with Carson's formula. One advantage of their method is the fact that it is very easy to assume different earth resistivities for each of the layers. Gary, Deri, et al. calculate self and mutual impedances with the simple formulas originally proposed by Dubanton [195], o Ho, 20hyt BD + jt 52 mn i-internal =, * *'i-internai? Coy 4-5 and Lip = jobs 1p OTR rea (4.4) in which P represents a complex depth, p= j/—e 4.5) 7 (ae: (as) All other parameters are explained after Eq. (4.8), except for x;, = horizontal distance between conductors i and k (Fig. 4.4), and‘f = earth resistivity. The results agree very closely with those obtained from Carson's formula, with the differences peaking at 9% in the frequency range between 100 Hz and 10 kHz and being lower elsewhere. This is a very good agreement, indeed, and Eq. (4.3) and (4.4) may therefore supplant Carson's formula some day. Fig. 4.2(b) shows a comparison of positive and zero sequence parameters of a typical 500 kV line. Wait [196] developed the complex depth formula apparently independently, at the same time as Dubanton. Its mathematical Verification, starting from Carson's formula, was shown by Deri and Tevan [197]. Alvarado and Betancourt [198]’ improved its accuracy by adding an extra (closed-form) term. Equations (4.3) and (4.4) can be used to explain why L'pg, = L'yir L'gyy or 1 to 2 hep) Loos = 5 In on [nape (h, a, x = geometric mean average height, direct distance, and horizontal @istance) is not completely constant, but decreases from 1.417 mH/mile at low frequencies to 1.37 mH/mile at very high frequencies in Fig. 4.2(b). For very low frequencies, f goes to infinity. Then Lies (ono) * Sa In(7) 4-5a At very high frequencies, PB goes to zero. Then Lp 5 ne since is less than 1.0, L',., decreases slightly. The same difference shows up in the positive sequence capacitance if the influence of height is ignored in the handbook formula Eq. (4.52). Fig. 4.2(a) - Alternative to Carson's formula: Ground represented as layers 1,2, ... n on's formula Carson's formula for homogeneous earth is normally accurate enough for power system studies, especially since the data for a more detailed multilayer earth return is seldom available. The Supporting routine CABLE CONSTANTS does have an option for multilayer or stratified earth, however. 4-6 Carson's formula is based on the following assumptions: (a) The conductors are perfectly horizontal above ground, and are long enough so that three-dimensional end effects can be neglected (this nakes the field problen two-dimensional). The sag is taken into account Bg o/ (A/nite) Rugs (a/mite) aot ae 10? 1 1 jt rete 107? io? 104 ao? 10? aot 104 17 f (iz) £ iz) Leero (mH/mile) ppg (nt/nile) 10 1.5, 5 1a CoN aot ro4 10” ro? 104 x07 £ (iz) £ (iz) Fig. 4.2(b) - Alternative to Carson's formula: formula by Gary, Deri et al. (comparison with Carson's formula for a typical 500 kV line with bundle conductors; skin effect in conductors ignored [199] 47 indirectly by using an average height above ground (Fig. 4.3). (b) The aerial space is homogeneous without loss, with permeability u, and permittivity ¢,- (c) The earth is homogeneous with uniform resistivity p, permeability p, and permittivity ¢,, and 1s bounded by a flat plane with infinite extent, to which the conductors are parallel. The earth behaves as a conductor, dees, 1/p >> weg, and hence the displacement currents may be neglected. Above the critical frequency fo.4¢4¢q) 7 1/(2ne,p), other formulas 135,36] must be used (for p = 10 000 Gm in rocky ground, fos tycg) * 1.8 Miz, which is still on the high side for most EMTP line models). (4) The spacing between conductors is at least one order of magnitude larger than the radius of the conductors, so that proximity effects (current distribution within one conductor influenced by current in an adjacent conductor) can be ignored. The conductor profile between towers (Fig. 4.3) can be described (a) as a parabola for spans < 500 m, (b) as a catenary for 500 < spans < 2000 m, and (©) as an elastic line for spans > 2000 m. Fig. 4.3 - Conductor profile between towers If the parabola is accurate enough, then the average height above ground is b= height at atdopen +} ass, 6.6) which is the formule used by both supporting routines LINE CONSTANTS and CABLE CONSTANTS. The elements of the series impedance matrix can then be calculated from the geometry of the tower configuration (Fig. 4.4) and from characteristics of the conductors. For the self impedance, Cheah BB yg) # So ge 8 SER tacernar + ED 7) od cee @ i-internal * Fig. 4.4 - Tower geometry and for the mutual impedance By OD, sgt mart 2 ag AK y axe Bryn Ryo aR + Me ze dn ge ae (4.8) with p, = permeability of free space. Using uyf2n = 20107 B/km (4.9) produces impedances in Q/km. The parameters in Eq. (4.7) and (4.8) are R's anternar 7 &€ Feststance of conductor 4 in Q/unte length, hy = average height above ground of conductor 4, Pay = distance between conductor 4 and image of conductor k, ay = distance between conductors 1 and k, ty = radius of conductor 4, Xt internal 7” internal reactance of conductor i, o = 2nf with £ = frequency in Hz, AR, Ax = Carson's correction terms for earth return effects. Carson's correction terms AR’ and AK’ in Eq. 4.7 and 4.8 account for the earth return effect, and are functions of the angle ¢ ($=0 for self tnpedance, ¢"94, in Fig. 4.4 for mutual impedance), and of the parameter a: a bwSe10 ee SE, (4.10) 49 with D = 2h, in m for self impedance, Dy, in m for mutual impedance, 9 = earth resistivity in Om. AR! and AX’ become zero for ar (case of very low earth resistivity). Carson gives an infinite integral for AR' and AX", which he developed into the sun of four infinite series for a <5. Rearranged for easier programming, it can be written as one series, and for impedances in Q/km, becones ant buet0™*{ AK'= 4ye10™*{ 3(0.6159315-1na) ~byascosy +byarcosd 2 2, 2 #b)[(cy-Ina)a"eos2g + 9 a7sin2g] =a ga7c024 +b 3070839 +b4a°cos36 ~a,a* cosh =p, [(c,-Ina)a“coste +) a4sinds] a a they 5, 5 “bea cos5p +b,a7cos5p 6 6 6 Ho gl(egrinayacoss9 + 9a°sin69] 4,2 %coa69 +b 7 c0876 +b2acos7p 8 8 8 Aga cos8> “byl (cg-Inaya®cos8e + ga°sin8»} Son) + ad in Q/km (4.11) Each 4 successive terms form a repetitive pattern. The coefficients by, cy and d, are constants, which can be precalculated and stored in lists. They are obtained from the recursive formulas: tga bya] reiSS, with the starting value by + by for even eubscripte, (4.12) aia = cup +7 + qpy with the starting value cy = 1.3659315, 1-2 4-10 with eign = #1 changing after each 4 successive terms (sign = +1 for 4 = 1,2,3,45 sign = -1 for 1 = 5,6,7,8 ete.). The trigonometric functions are calculated directly from the geometry, wy tM ae ging a BE cont and stn OE and for higher-order terms in the series from the recursive formulas atcosio = [a tcos(t-1)9 + cose - at tein(i-1) + sing] + a (4.13) ateinio = [a Teos(i-1)9 + sing + a* Tein(i-1)9 + cose] + a For power circuits at power frequency only few terms are needed in the infinite series of Eq. 4:11. However, at higher frequencies and for wider spacings (e.g., in interference calculations) more and more terms must be taken into account as the parameter a becomes larger and larger [37, discussion by Dommel]. Once Carson's series starts to converge, it does so fairly rapidly. How misleading the results can be with too few terms in the : If the series of Eq. 4.11 is illustrated for the case of a= 4 and ¢ = series were truncated after the Ist, 2nd, ..., 15th tera, the percent error in Re{2},} would be +312, -748, -16, 4798, -416, +365, -121, -93, 428, -15, 45.2, 41.7, -0.35, +0-14, -0.04. For a > 5 the following finite series [38] is best used: - Z sos , 20839 , 3eosSp _ 45c0874) , ha is ee eee tes re @ a @ a in g/km (4-14) ~ 20830 4 3eos5e , 45cos7¢) | hw + 1o™* Seep eee) eee a a a a Internal impedance and skin effect In the old days of slide-rule calculations, the internal reactance a 2h ; nd external reactance w a2 in < for ss earth were often Xiaernat ad external reactance w 72 tn 2" for losses earth were oft. 4-1 combined into one expression, by replacing radius r with the smaller “geometric mean radius” GYR to account for the internal magnetic field, rn u, Dr eiiee = Be gn 20 © tm = * Mavernat ~ ° Te GR GMR was often included in conductor tables. Instead of or in addition to (4-15) GMR, North American handbooks have also frequently given the “reactance at 1 * foot spacing’ ) Xe which is related to CMR, . Po. 1 (foot) 7 4 a GHRGFeaty ” with GMR in feet (or in m if Xj is to be the reactance at 1 m spacing). (4-16) The concept of geometric mean radius was originally developed for nonmagnetic conductors at power frequency where uneven current distribution (skin effect) can be ignored. In that case, its meaning is indeed purely geometric, with GYR being equal to the geometric mean distance among all elements on the conductor cross section area if this area were divided into an infinite number of equal, infinitesimally small elements. For a solid, round, nonmagnetic conductor at low frequency, our/e = 1/4 ‘This formula changes to maf GMR/r = & if the conductor is made of magnetic material with relative permeability us its geometric meaning is then lost. If skin effect is taken into account, its geometric meaning 1s lost as well. ‘The nane geouetric mean radius is therefore misleading, and it 1s questionable whether it should be retained. Eq. (4-15) gives the conversion formula between GMR and internal reactance, Done pane comes from the positive sequence reactance formula X',,,2u 5° an Gag discussed in Eq. (4.56), for the case where the spacin§ among the three phases (expressed as geometric mean distance GMD) is 1 foot, with GMR given in feet as well. 412 esr = The internal reactance can be calculated for certain types of conductors [39, (417) x Ho internat! ( 3) 40], as part of the internal impedance Rt Since ‘internal * internal’ Xinternal £8 OMly a very small part of the total reactance for nonmagnetic conductors, its accurate determination is somewhat academic. More important is the calculation of Ri taraay» because the increase of resistance with frequency due to skin effect can be considerable. The internal impedance of solid, round wires can be calculated with well-know skin effect formulas, with R} being of more practical ‘internal, Stranded conductors can usually be approximated as interest than X} se gonai* solid conductors of the same cross-sectional area”) [41]. It has been claimed that steel-reinforced aluminum cables (ACSR) can usually be approximated as tubular conductors when the influence of the steel core is negligible, which is more Likely to be the c: of aluminum strands, since the magnetization of the steel core caused by one e with an even number of layers layer spiralled in one direction is more or less cancelled by the next layer spiralled in the opposite direction. The supporting routine LINE CONSTANTS uses this approximation of an ACSR as a tubular conductor. If the magnetic material of the steel core is of influence, then calculations probably become unreliable, and current-dependent, measured values should be used instead. Since the solid conductor is a special case of the tubular conductor, the Supporting routine LINE CONSTANTS uses only the formula for the latter, which is described as Eq. (5.7b) im Section 5.1. Table 4.1 shows the increase in resistance and the decrease in internal inductance due to skin effect for a tubular conductor with R}=0.0398 o/mile, ratio inside radius/outside radius q/r = 0.2258 (Fig. 4.5), and p= 1.0. ‘The internal inductance of a tube at de is (48, p64} *)there are cases, however, where this approximation is not good enough. More accurate formulas are needed, for instance, for calculating the attenuation in power line carrier problems [39], as explained in Appendix VIL. The GMR of stranded conductors at low frequencies, where skin effect is negligible, is derived in Appendix VIII.2. 4-13, Table 4,1 - Skin effect in a tubular conductor fi) Bago © tacernat-ac!™'tnternal-de 2 1.0002 0.99992 4 1.0007 0.99970 6 1.0015 699932 8 10026 0.99879 10 Too 0199812 20 rotss 0199254 40 (1.0632 0.97125 60 ‘1.1347 0.93898 80 1.2233 0.89946 100 13213 85639 200 117983 0166032 400 2.4554 0.47004 600 2.9421 0.38503 800 3.3559 0.33418 1000 3.7213 0.29924 2000 5.1561 221204 4000 7.1876 0.15008 6000 8.7471 0.12258 8000 10.0622 0.10617 10000 1112209 0.08497 20000 15.7678 0.06717 49000 2211988 0108750, 60000 27.1337 0.03879 80000 512942 0.03359 100000 34.9597 0103008 200000 9.3413 0.02124 ‘ooc00 69.6802 0.01502 600000 85.2870 0.01227 800000 98.4441 01062 1000000 110.0357 0.00950 2000000 155.5154 0.00672 4000000 219.8336 0.00475 Fig. 4.5 - Tubular conductor Percent of Equal Current Distribution 4-14 2 » 22-1074 a E-3g-t i ee cto eee ear=r a or 0.454866 + 10-" H/km in this case. At high frequencies, Rivterng) ~ Xneernai’ With both components being proportional to yi. This is the region of pronounced skin effect. From Table 4.1 it can be seen that Rivegrag) and Xlnternal 2%¢ almost equal at 10 kiiz (difference 2.2%), with the difference decreasing to 0.7% at 100 kiiz, or 0.2% at 1 Miz. Subeonductor Position No. a %0 OF Decree of asymmetry: : t oo eta wl as we fron pie en measured fF - = - ~ computed rob 41 2 3 ‘ 5 ‘ 7_,s Subconductor Position No. Fig. 4.6 - Current distribution within an 8-conductor bundle [42] - © 1976 IEEE 415 Example for using series impedance matrix of individual conductors The matrix of Eq. (4.2) can be used to study the uneven current distribution within a bundle conductor. Fig. 4.6 shows measured and calculated values for the unequal current distribution in the 8 subconductors of an asymmetrical bundle for various degrees of asymmetry [42]. Asymmetrical bundling was proposed to reduce audible noise, but this advantage is offset by the unequal current distribution, The currents in this case were found from Eq. (4.2) with an 8 x 8 matrix, assuming equal voltage drops in the 8 conductors, [1] = (21 [av/dx) (4.18) 4141.2 Shunt Capacitance Matrix The voltages from the 13 conductors in Fig. 4.1 to ground are a function of the line charges: Pua Pia weePhyas | [a vy, PS BS Fs a aie ear 22 2,13 | |% ee is} [Pisy1 Pis2, st Pisyas} [913 with ay = charge per unit length on conductor 1, or in the general case (v] = (PN [al (4.19) Maxwell's potential coefficient matrix [P'] is real and symmetric. Its elements are easy to compute from the geouetry of the tower configuration and from the conductor radii if the following two assumptions are made: (a) the air is lossless and the earth is uniformly at zero potential, (b) the radii are at least an order of magnitude smaller than the distances among the conductors. Both assumptions are reasonable for overhead lines. Then the diagonal element becomes 4-16 Pi Be dn a (4.20) and the off-diagonal element Mery 1, hue Pe Rat me ae (4.21) with ©, = permittivity of free space. The factor 1/(2re,) in these equations ts c+ "2, where c te the speed of Light. With © = 299 792.5 ka/e and 4, / Qn) wa + 10- wea, 18 follows that 1/(2ne,) = 17.975109 + 10° kn/F (4.22) The inverse relationship of Eq. (4-19) yields the shunt capacitance matrix [C'], [a] = (e'}{v}, with fot} = ppry7t (4.23) The supporting routine LINE CONSTANTS uses a version of the Gauss-Jordan process for this matrix inversion which takes advantage of symmetry [43]. This process was chosen because it can easily be modified to handle matrix reductions as well, which are needed for eliminating ground wires and for bundling conductors. Appendix III explains this Gauss-Jordan process in more detail. The capacitance matrix [C'] is in nodal form. This means that the diagonal element C}, is the sum of the shunt capacitances per unit length between conductor { and all other conductors as well as ground, and the off- diagonal element Cj, = Ci, is the negative shunt capacitance per unit length between conductors i and k. An example for a three-phase circuit from [44, p+ 457] 4s shown in Fig. 4.7, with 12.161 = 26625 = 2.625 [ct] = ]- 2.625 11.729 ~ 1,349] nF/mile - 2.625 - 1.349 11.729 SF Cip-mutuat ~ 2*625> Chs-mutuat ~ 2-625 C3 mutuar ~ 1349 Cr-ground ~ 7.755 nF/mile, etc. 4-17 1.349 nF/mile 10 feet _10 feet 2.625 nF/mile be ae Z N I - “i 20 19-1385 obtained from the | ! i following data: 40 feet conductor diameter | = 0.5 inches 6.911] nF/mile Fig. 4.7 - Mutual and shunt capacitances For ac steady-state conditions, the vector of charges (as phasor values) 4e related to the vector of leakage currents [- $4] by =-1 ah ta = -3 GH (4.24) Therefore, the second system of differential equations 1s ~ 1G) = jo (or v) (4.25) which, together with Eq. (4.2), completely describes the ac steady-state behaviour of the multi-conductor line. Shunt conductances G' have been ignored in Eq. (4.25), because their influence is negligible on overhead lines, except at very low frequencies approaching de, where the line behaviour is determined by R! and G', with ul’ and uC becoming negligibly small. With G', the complete equation is - 8 = wiv (4.264) where (y'] = [6] + jw [e"} (4.260) At very high frequencies, the shunt capacitances are algo influenced by 4-18 earth conduction effects, and correction terms must then be added to Eq. (4.20) and (4.21). However, the earth conduction effect is normally negligible beloy 200 kHz to 1 MHz (45]. In that case, the capacitances are constant, in contrast to series resistances and series inductances which are functions of frequency. 4.1.1.3 Insulated overhead conductors If the overhead conductor has a layer of insulation around it, but no sheath on the outside, this insulation layer can be ignored in the calculation of the series impedance matrix, because the relative permeability 4, of insulation and air are identical. In the capacitance calculation, the insulation can be taken into account by using a modified radius in Eq. (4-20), : (tin Senoditied * Four |Z ' where x, = conductor radius (inside radius of insulation) and rq, = outside radius of insulation, ¢, = relative permittivity of insulation. This formula is derived from the capacitance of a single conductor, which can be obtained by integrating the electric field intensity from infinity to r,. 4.1.2 Line Parameters f: valent Conductore Equations (4.2) and (4.19) for all individual conductors contain more information than is usually needed. Generally, only the phase quantities are of interest. For the case of Fig. 4.1, the reduction from 13 equations to 6 equations for the phases R,S,T,U,V,W is accomplished by introducing the following conditions, for grounding conductor 13: (Vj) /dx = 0 in (4.2), vy = 0 in (4.19), for bundling conductors 1 and 2 into pha: av, dv, dv, ees oer ers in (4.2), ana 4a ae v= ye ve in (4.29) and analogous for bundling the other phases. With these conditions, the matrices gan be reduced to 6 x 6, as explained next. These reduced matrices will be called matrices for the equivalent phase conductors. 4.1.2.1 Elimination of Ground Wires Normally, ground wires are continuous and grounded at every tower”, which are typically 250 to 350 m apart. In that case it is permissible for frequencies up 0 approximately 250 kHz to assume that the ground wire potential is continuously zero [46]. This allows a reduction in the order of the [Z']~ and (P"]-matrices, with the reduction procedure being the same for *)Non-continuous " egmented” ground wires are discussed in Section 4.1.2.5 4-19 both. Let the matrices and vectors in Eq. (4.2) be partitioned for the set “u” of ungrounded conductors, and for the set “g” of ground wires, av, /dx Zig) 2g) - = (4.27) lav, /ax] (25) (23,1 (1 Since [V,] and [4V,/dx] are zero, Eq. (4.27) can be reduced by eliminating (1,1, = eeaucea! ty] (4.28a) where . =) = tecates te U2egucea! ~ (Zyl - 2ygHl2p 1123) (4.286) Rather than using straightforward matrix inversion and matrix multiplications in Eq. (4.28), the more efficient Gauss-Jordan reduction process of Appendix III is used in the supporting routine LINE CONSTANTS. [P'] is reduced in the same way, and [C!.4,ceg] 18 found by inverting [Plegucea!+ may appear as if less work were involved in reducing [C'], where the At first sight it reduction simply consists of “scratching out” the rows and columns for ground wires However, [C'] must first be found from the inversion of [P"], and it is faster to reduce a matrix than to invert it. 4.1.2.2 Bundling of Conductors on high voltage power lines, bundle conductors are frequently used, where each “phase” or bundle conductor consists of two or more subconductors held together by spacers every 100m or so. The bundle is usually symnetrical ($ = 1.0 in Fig. 4.6), but unsymmetrical bundles have been proposed as well. ‘Two methods can be used for calculating the line paraneters of bundle conductor: With the first method, the parameters are originally calculated with each subconductor being represented as an individual conductor. Since the voltages are equal for the subconductors within a bundle, thie voltage equality is then used to reduce the order of the matrices to the number of “equivalent phase conductors”. With the second 4-20 method, the concept of geometric mean distances is used to replace the bundle of subconductors by a single equivalent conductor. Both methods can be used with the supporting routine LINE CONSTANTS. The supporting routine CABLE CONSTANTS is limited to the second method. Method 1 ~ Bundling of subconductors by matrix reduction As in the elimination of ground wires, the matrix reduction process is the same for [Z'] and [P"], and will therefore only be explained for [Z"]. Let us assume that the individual conductors i, k, 1, m are to be bundled to make up phase R. Then the conditions +h titty = Ig and av, av, aVy eV ag “dx "dk " “dx " “dx “ “Gk” must be introduced into Eq. (4.2). The first step is to get I, into the equations. This is done by writing Ip in place of I,. By doing this, an error is of course made, which amounts to the addition of terms 2( + 1, + Ty) in all rows 4; they must obviously be subtracted again to keep the equations correct, In effect, this means subtraction of column i from columns k, £,m. These changes are shaded in Fig. 4.8, ‘ kee subtract colum 4 from colums ky by mi colum i remains wmchangea Fig, 4.8 - First step in bundling procedure 4-21 Columns k, £m are assumed to be the last ones in the matrix to make the explanation easier. The currents I,, I,, I, are still in the equations after execution of the first step of Fig. 4.8. To be able to eliminate them, there should be zeros in the lefthand side of the respective rows, This is easily accomplished by subtracting row i from rows k, £, m, which produces zeros because 4V,/dx = dV,/dx etc. These changes are shaded in Fig. 4.9. The equations are now in a form which permits elimination of I,, I,, In, in the same way as elimination of ground wires in Eq. (4.28). The four rows and columns for subconductors 1, k, 2, m are thereby reduced to a single row and column for bundle conductor R. Method 1 ie more general than method 2 discussed next. For instance, it can easily handle the unequal current distribution in asymmetrical bundles described in Fig. 4.6. subtract row {from rows ky tym row a remains unchanged Fig. 4.9 - Second step in bundling procedure Method 2 — Replacing bundled subconductors with equivalent single conductor This method was developed for hand calculations [47], and while theoretically not limited to symmetrical bundles, formulas have usually only been derived for the more important case of symmetrical bundles. The following formulas are based on the assumption that (a) the bundle is symmetrical ($ = 1.0 in Fig. 4.6), and (b) the current distribution among the individual subconductors within a bundle is uniform. Fig. 4,10 - Symmetrical bundle with N individual eubconductors With these assumption, the bundle can be treated as a single equivalent conductor in Eq. (4.15) by replacing GYR with the equivalent geometric mean radius of the bundle, *) Kom , (4.29) where GMR = geometric mean radius of individual subconductor in bundle, A= radius of bundle (Fig. 4.10). Similarly, the radius r in Eq. (4.20) must be replaced with the equivalent radius ‘equiv Fee (4.30) Comparison between methods 1 and 2 Both methods for bundling conductors give practically identical answers, at least in the example chosen for this comparison. The example was a 500 kV three-phase line with horizontal tower configuration, with phases 40 feet apart at an average height above ground of 50 feet. The symmetrical bundle consisted of 4 subconductors spaced 18 inches apart. Conductor diameter = 0.9 inches, de resistance = 0.1686 9/mile, GMR = 0.3672 inches, r, 7.80524 inches from Eq. (4.30), and GMR, ‘equiv (4.29). Table 4.2 compares the results in the form of positive and zero ‘equiv ~ = 7.41838 inches from Eq. sequence parameters at 60 Hz. Obviously, the results are practically identical. ¥ ) Equation (4.29) is derived in detail in Appendix VIII.1. 4-23 Table 4.2 - Comparison between methods 1 and 2 for bundling Positive and zero [Method 1 (Bundling by|Method 2 (Equivalent sequence parameters| matrix reduction) conductors) ‘at 60 Hz Rigg Wate) 0.042223, 0.042205 xP08 ( W/mile) 0153394 0153399 cre wF/mte) 0.021399 0.021397 Beero( Wate) 0.31740 0.31738 xfopo( Wate) 2.0065, 210065 Crore uF/mile) 0.013456 01013455, 4.1.2.3 Reduced Matrices for Equivalent Phase Conductors For the case of Fig. 4.1, elimination of ground wires and bundling of subconductors reduces the 13 x 13 matrices for the individual conductors to 6 x 6 matrices for the phases, e.g., for the series impedances, ty Fy ty ty ty Av : Xy My I 1phase! (4.31) and 424 aly, ~ (BBP) = jalcr (4.32) phase! (phase! For a three-phase single circuit with phases A, B, C, Eq. (4.31) would have the form av A e ogr gt a an ap Zac] [Ta avy -|-]- Ze el |, (4.33) ay, c . a %e5 ec} [To The diagonal element 2}, in Eq. (4.33) is the series self impedance of phase k for the loop formed by phase k with return through ground and ground wires, and the off-diagonal element 2}, is the series mutual impedance between phases i and k. The self impedance of phase k is not the positive sequence impedance. To obtain impedances which come close to the positive sequence values, we would have to assume symmetrical currents in Eq. (4.33), 2 Bhizoe I, = aI, and I, = al,, with a= @)120°, and then express the voltage drop in phase A as @ function of I, only, ay, Gx 7 7 a-syam Tar ME A sym 1 Qo + 7 (2, + a2, + ath, (4.34a) and similarly for phases B and C, 2" seyam Tpr MECH 2 oom + 2a (hy + a2h, + a7Zh), (4.340) ~ > c-syam The values of the three impedances Z\ aos Zi symm? 2-gymm 18 Ede (4-34) ate not exactly equal, but their average value is the positive sequence Igy with 2; 228 . oesyan 7 (Bbq + A Zhe + a2ho). (4.34¢) impedance. Because of slight differences in the three values, the voltage drops are slightly unsymmetrical (or the currents become slightly unsymmetrical for given symmetrical voltage drops). As discussed in Section 4-25 4.1.3, transposing a line eliminates or reduces these unsymmetries at power frequency, though not necessarily at higher frequencies. In the capacitance matrix of a three-phase line, Ct, would be the sum of the coupling capacitances to phases B and C and of the capacitance to ground, and Cl, would be the negative value of the coupling capacitance between phases A and B. Assuming symmetrical voltages, Eq. (4.32) would show slight unsymmetry in [dI., ,.,/dx], analogous to that of Eq. (4.34). phase’ 4.1.2.4 Nominal Circuit for Equivalent Phase Conductors The matrices in Eq. (4.31) and (4.32) are the basis for practically all EMTP line models. Even in studies where ground wires must be retained, it is still these matrices which are used, with phase numbers assigned to the ground wires as well. A three-phase line with one ground wire is conceptually a four-phase line, with phase no. 1,2,3 for phase conductors A,B,C and phase no. 4 for the ground wire. One type of Line representation uses cascade connections of nominal weircuits, as discussed in Sections 4.2.1.1 and 4.2.2.1. This polyphase nominal weircuit with a series impedance matrix and equal shunt capacitance matrices at both ends, as shown in Fig. 3.10, 1s directly obtained from the matrices in Eq. (4.31) and (4.32), [R) + jel) = 2+ (2g) (4.35) and 4 jutel = 4 jut (cr,,.0 (4.36) z z hase B where 21s the length of the line. The cascade connection of nominal circuits approximates the even distribution of the line parameters reasonably well up to a certain frequency. It does ignore the frequency dependence of the resistances and inductances per unit length, however, and is therefore reasonably accurate only within a certain frequency range. Strictly speaking, it may not be quite correct to treat the real part of [ Urage! a8 2 resistance, and the Imaginary part as a reactance, as done in 4-26 Bq. (4.35), especially for lines with ground wires. For a three-phase line with phases A,B,C and ground wire g, the original 4 x 4~matrix is reduced to a3 x 3-matrix with elements 2 ce . an _ Zig-original “kg-original Zy-reduced ~ “ik-original 4.37) Zag-original Even tf Zi, riginar Could be separated into resistance and reactance without any doubt, the real part of the second term in Eq. (4.37) depends on the imaginary parts of the three impedances as well, unless the R/X-ratios of all three impedances were equal. There is also some doubt about separating Zip-ortginal into resistance and reactance because of the earth as an implied return conductor, as mentioned in Section 3.1, Nontheless, experience has shown that nominal reircuits do give reasonable answers in many cases, and they are at least correct at the frequency at which the matrices were calculated (and probably reasonably accurate in a frequency range around that specific frequency). Example for using nominal rcircuits Electrostatic and magnetic coupling effects from energized power lines to parallel objects, such as fences or deenergized power lines, are impor- tant safety issues, and have been well described in two IEEE Committee Reports [37,49]. A case of a fence running parallel to a power line (Fig. 4.11) is discussed here, as an application example for nominal weircuits.”) “Yor electrically short lines, as ia this example, electrostatic coupling effects can be solved by themselves with [Ctyagel, and magnetic coupling effects by thenselves with [Z%pagel+ For solving such cases with the EMTP, it 1s usually easier to use nominal rcircuits which combine both effects. With that approach, electrically long lines can be studied as well, provided an appropriate number of rcircuite are connected in cascade. 4-27 Power line conductors: 3.048 m 3.048 m Rinternal = 9-348 O/km ho Xj = 0.4755 Q/km (60 Hz) a h 2 PF (reactance at 1 m spacing) Giemeter = 12.7 nm | frequency = 60 Hz | | 12.192 m — 0 Fenc! Rinternal * 1-802 Q/km 3.048 m jg 9.244 solid conductor (nonmagnetic) ¥ diameter = 4.064 mm length = 2 km Fig. 4.11 - Fence 4 running parallel with power line phase conductors 1,2,3 By simply treating the fence as a fourth phase conductor, the following series impedance and shunt capacitance matrices are obtained: (0..4054+50.9859 symmetric! 0.05744j0.4265 0.4054+0.9859 [2tnase!™ a/kan }0,0574440.4265 0.0574450.3742 0.4054+j0.9859 10.0581+j0.3168 0.0581+j0.3291 0,0581+j0.3044 1.8607+j0.9953 and 7.5709 symmetric 1.6266 7.3088 ‘prasel"| 1 6304 0.0049 7.2999 ie 0.1688 0.2758 += -0.1189 «6.9727, From these matrices, the nominal n-circuit matrices are calculated with Bq. (4.35) and (4.36). Assume that the fence is insulated from the posts and nowhere grounded. To find the voltage on the fence due to capacitive coupling, simply connect voltage sources to phases 1,2,3 at the sending end, and leave 1,2,3 at the receiving end as well as 4 at both ends open-ended. Assuming V = 345 kV RMS, line-to-line, the fence voltage becomes V, = 3.97 kV. If 4-28 phase 1 were at zero potential because of a phase-to-ground fault, with phases 2 and 3 still at rated voltage 345//3 kV, then the fence voltage would increase to V, = 6.84 kV. These answers are practically independent of fence length. Now assume that the 2 km long fence is grounded at the sending end and open-ended at the receiving end. To find the voltage in the fence for a load current of 1 kA RMS, simply add current sources to phases 1,2,3 at the receiving end, with symmetrical voltage sources at the sending end. Phase 4 is connected to ground at the sending end and open-ended at the receiving end. The answer will be V, racetving end™0*043 KV» which increases dramatically to 6.442 kV if the currents are changed to I,=10 kA, I,=I,-0 to simulate a phase-to-ground fault. For this last case, the fence current would be 1.526 kA if the fence were grounded at both ends. These answers are practically independent of the voltage on phases 1,2,3, which can easily be verified by setting them zero. 4.1.2.5 Continuous and Segmented Ground Wires (a) Circulating Currents in Continuous Ground Wires Assume that ground wire no, 13 of Fig. 4.1 is grounded at each tower. If the ground wire is not eliminated, then the series impedance matrix for equivalent phase conductors will be a 7 x 7 matrix. Its elements can then be used to calculate the longitudinally induced voltage in the ground wire, av, -—8en . . . Ge Ugly t Mhglg t ++ Bayly t Ugly (4.38) If tower and tower footing resistances are ignored, then V at all towers as long as span << wavelength, or e . Bale wy 7 (4.39) Since the mutual impedances from the phase conductors to the ground wire are never exactly equal, the numerator in Eq. (4-39) does not add up to zero even if the phase currents are symmetrical. Therefore, there is a nonzero ground wire current I,, produced by positive sequence currents, which circulates 4-29 through ground wire, towers and ground (Fig. 4.12). This circulating current produces additional losses, which show up as an increase in the value of the positive sequence resistance, compared with the resistance of the phase conductors. Handbook formulas would not contain this increase, but the elimination of the ground wires discussed in Section 4.1.2.1 will produce it automatically. In one particular case of a single-circuit 500 kV line, this increase was 6.5%. 9 Fig. 4.12 - Circulating current in ground wire The inclusion of tower and tower footing resistances may change the results of Eq. (4.39) somewhat. If we assume equal resistance at all towers, then it appears that the voltage drop produced by the current in the left loop (Fig. 4.13) 4s cancelled by the voltage drop produced by the current in the middle loop, and Eq. (4.39) should therefore still be correct, except in the very first and very last span of the line. This assumes that the phase currents do not change from one span to the next, which is reasonable up to a certain frequency. Fig. 4.13 - Cascade connection of loops 4-30 (b) Segmented Ground Wires To avoid the losses associated with these circulating currents, some utility companies use segmented ground wires which are grounded at one tower, and insulated at adjacent towers to both ends of the segmentation interval, Where they are interrupted as well (Fig. 4.14). "7-configuration" in segmentation interval ‘4 = insulator Fig. 4.14 - Segmented ground wires They still act as electrostatic shields for lightning protection, but when struck by lightning, the segmentation gaps and the small insulators will flash over, thereby making the ground wire continuous again. The supporting routine LINE CONSTANTS has an option for segmented ground wires, viich ignores.’ them in the calculation of the series impedance matrix since they have no influence on the voltage drops in the phase conductors, but takes them into account in the calculation of the capacitance matrix because the electrostatic field is not influenced by segmentation. (c) Reduction Effect of Continuous Ground Wires on Interference Interference from power lines in parallel telephone lines becomes a problem if there are high zero-sequence currents in the power line, e.g., in case of a single-phase-to-ground fault. Assume a three-phase line with one ground wire g and a parallel telephone line P as shown in Fig. 4.15. For Zero sequence currents, which implies equal currents in phases A,B,C, the voltages in P induced by currents in A,B,C will add up in the same direction "an exception are studies where it can be assumed that the gaps and insulators have flashed over. For such studies, ground wires must be treated as continuous, as suggested by W.A. Lewis. Switching and lightning surge studies may fall into this category. 4-31 (Fig. 4.16). The voltage induced from the ground wire current I, will have Fig. 4.15 - Parallel telephone line P close to a power line with phases A, B, C and ground wire g opposite polarity, however, since this current flows in opposite direction, thereby reducing the total induced voltage -@V,.inmceu/OX. Part of this beneficial reduction may be offset by an increase in the zero sequence currents because wav,/ax OSS Beata 2 ppta Pi pcte Fig. 4.16 - Induced voltage caused by currents 1, = I, = Iq and by I ground wires also reduce the zero sequence impedance of lines (typically by 5 to 15% with one steel ground wire, or 15 to 30% with one ACSR ground wire). The reduction effect of the ground wire on interference can be included in the calculations in two different ways: (a) Obtain the mutual impedances from watrices in which ground wires have been eliminated and in which the parallel telephone line has been retained as an additional conductor. Then the reduction effect of the ground wires 1s automatically contained in calculating the magnetically induced voltage from Vp nduced ae and, if needed, the electrostatically induced voltage for an insulated At va Ja Zha-redueedl + Zpp-reduced!B + Zpc-reduoed!C (4-408) parallel telephone line from (b) 4-32 Vp (4-40b) o Vad -reduced’c* °pp-reduced”P Sba-reduced”A* Sbp-reduced”s* “1 Calculate the mutual impedances from P to the phases as well as to the ground wires (or obtain them from matrices in which the ground wires were retained), and recover the value of the ground wire currents with a “screening matrix” from the phase currents. By setting V,=0 in Eq. (4.27), the ground wire currents are obtained Ul ~ (23) 12 ty (4.41) + "Fecreen! with “u” indicating ungrounded phase currents here. The screening ‘screen! 18 the transpose of the distribution factor matrix [D2] of Eq. (IIT.14) 4m Appendix ITT, and as indicated there, can matrix [F. easily be obtained as a by-product of the matrix reduction process. As an example, Fig. 4.17 shows the standing waves of the phase currents of 80 8 LINE CURRENTS IN AMPERES > 3 > >< i: AAA ~ — | — Tt L n 5 | | —t | 4 PN LE XIN MILES (DISTANCE FROM SENDING ENO) Fig. 4.17 - Currents of sixth harmonic in HVDC line [11]. © 1969 IEEE 4-33 the sixth harmonic of 60 Hz in the two poles A,B of the Pacific Intertie HVDC line, as well as the currents in the two ground wires recovered with Eq. (4.41) [11]. 4.1.3 * ve_and_Zero Sequence Parameters of Balanced”) Lines A “balanced” transmission line shall be defined as a line where all @tagonal elements of [2/,,5¢) and [Cingge! are equal among themselves, and all off-diagonal elements are equal among thenselves, (4.42) Fig. 4.18 - Bipolar de line The only line which 1s truly balanced is the symmetric bipolar dc line (Fig. 4,18), where Z},=Z}0"Z! and Zj,=Z\. Single-circuit three-phase lines become more or less balanced if the line is transposed, as shown in Fig. 4.19, provided the length of the “barrel” (=3 sections, or one cycle of the transposition scheme) is much less than the wavelength of the frequencies involved in the particular study. While the Westinghouse Reference Book (51, “Yatso called “continuously transposed” in the EMTP Rule Book. 4-34 p 777] mentions that a barrel may be 80 to 160 km in length on long lines, a German handbook [52, p. 555] recommends that one barrel be no longer than 80 km (at 50 Hz, or 67 km at 60 Hz) for lines with triangular conductor configuration, or 40 km (at 50 Hz, or 33 km at 60 Hz) for other conductor configurations. Whatever the length of the barrel, it is important to z Ir qr Fig. 4.19 - Transposition scheme for single three-phase circuit realize that while a line may be reasonably balanced at power frequency, there may be enough unbalance at higher frequencies”. If the barrel length is much shorter than the wavelength, then series impedances can be averaged by themselves through the three sections, and shunt capacitances can be averaged by themselves, e.g., for the impedances of the line in Fig. 4.19, Bis Fhe Fim] Pe em Zi] [am Za Zak]] [#5 2m Zn] i gt gt n ogt gt 1 ogt as = lero a 3 ics Ace Mem) * [7m 7mm mi} * |7im 74s Air en 5 | Zot “ak Yom] [te tm 71a] [Bem “et Zi! Za ta 76) , | (4.43) with talc 1 1 23 (2, + y+ 2) atop + + gochett The averaging process for the shunt capacitances is analogous. "Yat the time of writing, studies at B.C. Hydro seem to indicate that transposed single-circuit lines with horizontal conductor configuration cannot be treated as balanced lines in switching surge studies. 4-35 4.1.3.1 Positive and Zero Sequence Parameters of Single-Circuit Three- Phase Lines Balanced single-circuit three-phase lines can be studied much easier with symmetrical or a, g, O-components because the three coupled equations in the phase domain, wom a =f 2) yaa (4.84 % m8 become three decoupled equations with symmetrical components, ~ Were! * Zerotzero - k= 2h . AV p96/8® = hog Tyog (4.45) - dx = 21 I, neg! ‘pos “neg or with a, 8, O-components, = 8 pc0/BE * Bepclaero wav jéx = 2, 1 (4.48) ~ aves = 2 d Zpos “8 Since transformation to symmetrical components involves complex coefficients, symmetrical components are not well suited for transient analysis where all variables are real, and are therefore only briefly discussed in Section 4.1.4, ‘The impedances needed in both systems (4.45) and (4.46) are the sane, however, namely Z},,, and 2),,. The balanced distributed-parameter line models in the EMTP use transformations to a, 8, 0- components, due to Edith Clarke [44], 4-36 1 Moag! ~ (7) Wphase! and (4.47) m1 Uy ggl © (TI pase! 0 where [vgagl = |v} > 1 & oO to4not with [7] 1 A} wm -4 -2] (ass) mz 1 0 8 -B ieee The columns in [T] and [T]”" are normalized; in that case [T] is orthogonal, try? = int (4.49) applying this transformation to Eq. (4.44) produces avi/ax] [ay + 224 0 07 fe ~ fav fax} = | 0 zh Zt o | |r). av o/dx| ° oa at} |r which is identical with Eq. (4.46), with ater siete ozs (4.50) Zoe" 25 ~ 2a (4.50b) Eq. (4.50) and its inverse relationship is the same as discussed previously 4-37 in Eq. (3.6) and (3.4), Going from the three coupled equations in (4.44) to the three decoupled equations in (4.46) allows us to solve the line as if it consisted of three single-phase lines, which is much simpler than trying to solve the equations of a three-phase line. The positive sequence inductance of overhead lines is practically constant, while the positive sequence resistance remains more or less constant until skin effect in the conductors becomes noticeable, as shown in Fig. 4.20. Zero sequence inductance and resistance are very much frequency- dependent, due to skin effects in the earth return. : conductor for akin effect oe calculation ca Fig. 4.20 - Positive and zero sequence resistance and inductance of a three-phase line 4-38 The shunt capacitance matrix of a balanced three-phase line becomes diagonal in a, g, O-components as well, with Choro 7 Oy + 208 (4.51a) Choe =~ Oh (4,51b) Which is the inverse relationship of Eq. (3.13). The capacitances are constant over the frequency range of interest to power engineers. Comparison with results from handbook formu: The positive and zero sequence parameters obtained from the supporting routines LINE CONSTANTS and CABLE CONSTANTS may differ from those obtained with handbook formulas. Since some EMIP users may make comparisons, it may be worthwhile to explain the major differences for a specific example. Assume a typical 500 kV line with horizontal phase configuration, with phases 40 feet apart at an average height above ground of 50 feet. Each phase consists of a synmetrical bundle with 4 subconductors spaced 18 inches apart. Subconductor diameter = 0.9 inches, de resistance = 0.1686 O/mile, GMR = 0.3672 inches. Throughout this comparison, the bundle conductors are ‘equiv 7 7+80524 inches from Eq. = 7.41838 inches from Eq. (4-29)+ represented as equivalent conductors with r, (4,30) and CHR gusy For positive sequence capacitance, most handbooks give the formula 2ne, Cea (4.52) ‘equiv ———— with d= “/d,54,cdy¢ (geometric mean distance anong the three phases). ‘This produces a value approx. 4% lower than the more accurate value from Eq. (4.51) for the 500 kV line described above. The formula for zero sequence capacitance in [52] and [53], 2ne, ‘zero 7 Z (Siemens) Pn (4.53) 4-39 with hl = Sarge (geometric mean height), = 3yp—BTE(weomet ric mean distance between one phase and image Pa ‘AB AC BC of another phase), can be derived by averaging the diagonal and off-diagonal elements in the (es phase has this averaging process implied in the [C’, J-matrix among themselves to account for transposition. Eq. (4.51) nase! Bateix. Both give practically the same answer, with results from Eq. (4-53) 0.23% lower than those from Eq. (4.51). In [51], Eq. (4-53) is further simplified by assuming Dy * 2h, 2ne, Stero™ (2h)? Cece (Westinghouse) mm Tequiv'n which produces a value 4% higher than the value from Eq. (4.51). While Eq. (4.54) 48 theoretically less accurate, the value may actually be closer to measured values because the influence of towers, which is neglected in all formulas, typically increases the calculated zero sequence capacitance by about 8 to 9% on 110 kV lines, about 6% on 220 and 380 kV lines, and about 4% on 700 kV lines [54, p. 218]. The formulas for zero and positive sequence impedances in most handbooks are based on the assumption that parameter a in Eq. (4.10) 1s so small that only the first term in the series of Eq. (4.11) must be retained. For normal phase spacings this is probably a reasonable assumption at power frequency 50 or 60 Hz. Then, after all diagonal and off-diagonal elements have been averaged out among themselves through transposition, se agro weed ARs aR z in ka, ' 4 ia and AX! = 2u+10™[0,6159315 ~ mn(2h,k "£)) tn a/km % (4.55) axt, * 2ue10™*[0,6159315 - m(D,k i in Vko ° with k= Gn 10 4-40 This leads to the expression 4 0 Gin /km (4.56) ‘equiv 20g = RL, + i2uel0 pos ~ Bac with Ri sac resistance of equivalent phase conductor. It is interesting that the influence of ground resistivity and of conductor height, which is present in 25 and 21, completely disappears here in taking the difference, 2) ,.92/- 2h. Eq. (4.56) is the formula found in most handbooks. Table 4.3 compares results from Eq. (4.50) with results from Eq. (4,56) for the 500 kV line described above with the following additional assumptions: Earth resistivity = 100 ma; skin effect within conductors ignored to limit differences to influence of earth return (that is, RL, = Ri, and GYR,,4, = constant). Table 4.3 - Accurate and approximate positive sequence resistance and inductance ACCURATE, APPROXIMATE and Lig, from Eq. (4.50)|Rigg and Lig from Eq. (4.56) £ Rt ao Rt u (iz) (mile) (ali/mile) (Wate) (ali/mile) 10-6 | 0.04215 1.417 0.04215 1.417 10 0.04215 1.416 0.04215 1.417 100 0.04229 1.416 0.04215 1.417 1 000 0.05003 1.416 0.04215 1.417 10 000 0.3528 1.413 0.04215 1.417 100 000 6.229 1.401 0.04215 1.417 Table 4,3 shows that L1,, from Eq. (4.56) is quite accurate over a wide frequency range, whereas Ri, increases (0.33% error at 100 Hz, but wrong by orders of magnitude at g becomes less accurate the frequency 100 kliz). The increase in RI,, in the higher frequency range 4s caused by eddy currents in the earth, as indicated in Fig. 4.21 for a bipolar de line Ground wires also influence the positive sequence impedance, as mentioned in Section 4.1.2.5(a). Both influences are ignored in Eq. (4.56), but automatically included in the method described here. 441 Fig. 4.21 - Eddy currents in earth The zero sequence impedance obtained from Eq. (4.55) is a . 658.87 2 . . 310 E Boro” ye * MGA) + Sowac™® an in km (4,57) 2 MReguty “n with £ in Hz, pin Mm, and GMR,,,, and d, in m, Eq. (4.57) i the sane equation as in [51, 52, 53]. Table 4.4 compares the approximate results from Eq. (4.57) with the accurate results from Eq. (4.50). The inductance Lt, ts reasonably accurate over a wide frequency range (-0.75% error at 100 Hz, -33% error at 100 Kitz), but the resistance Rt... is less accurate (4.6% error at 100 Hz, 159% error at 100 kHz). Table 4.4 - Accurate and approximate zero sequence resistance and inductance ACCURATE, APPROXIMATE Reero and Lerg from Eq. (4.50)|Rbero and Lig, from Eq. (4.57) f R ut R Lt (iz) | (smile) (mli/mile) (Waite) (ak/mi le) 10-6] 0.04215 13.94 0.04215 13.94 10 0.08905 6.170 0.08980 6.158 100 0.4960 5.084 0.5187 5.046 1 000 4.169 4.052 4.807 3.934 10 000 | 32.12 3.164 47.69 2.823 100 000 | 184.0 2.568 476.6 7 4-42 4.1.3.2 Positive and Zero Sequence Parameters of Balanced N-Phase Lines The EMTP can handle balanced distributed-parameter lines not only for the case of a three-phase Line, but for any number of phases M. For this general case, the a, §, O-tranaformation of Eq. (4.47) has been generalized to M phases, with the transformation matrix [55] Seth col. + = oS Lu MH 2 8 Lh-L L 1 am VID in-}L 0 -2 (4.58) va 3 oo c J-th 5 : row : ° + ° ° ° - os YORI) where again ~ t (17 = (7) (4.59) IT] of Eq. (4.48) is a special case of Eq (4-58) for M= 3 if we assume that the phases are numbered 2, 3, 1 in Eq. (4.47) and if thea, 8, O-quantities are ordered 0, f, -« (sign reversal on ¢). 4-43, * applying this M-phase a, B, O-transformation”? to the matrices of M-phase balanced lines produces diagonal matrices of the form ‘pos a ‘pos. a ‘pos with the first diagonal element being the zero sequence (ground mode) impedance, and the next M-1 diagonal elements being the positive sequence (aerial mode) impedance, Zerg 7 Zt (DZ, (4.60a) Bog 7 257 tp (4.600) and similarly for the capacitances, Cheeg yt RDC (4.61a) clooect-ct. (Pos ae ea To refer to the two distinct diagonal elements as zero and positive (4.61) sequence may be confusing, because the concept of sequence values has primarily been used for three-phase lines. “Ground mode” and “aerial mode” may be more appropriate. Confusion 1s most likely to arise for double circuit three-phase lines, where each three-phase line has its own zero and positive sequence values defined by Eq. (4-50) and (4.51) with symmetrical components used for each three-phase circuit, while in the context of this section the double-circuit line is treated as a six-phase line with different *)In the UBC EMTP, and in older versions of the BPA ENTP, Karrenbauer's transformation [57] is used instead, which produces the same diagonal matrices, but does not have the property of Eq. (4.59). This property is important because it makes the balanced line just a special case of the untransposed line discussed in Section 4.1.5. 4 zero and positive sequence values defined by Eq. (4.60) and (4.61). ‘The fact that the terms zero and positive sequence are used for M#3 as well comes from the generalization of symmetrical components of Section 4.1.4 to M phases with the transformation matrix [56, p. 155] [Sy-phase! * (4.628) with 5 al Ge exe 3 ZF G-1)Ge1) } (4.620) A spectal case of interest for symmetric bipolar de lines”) is M= 2 In this case [T] of Eq. (4.58) and [S] of Eq. (4.62a) are identical, t )-4 ‘ ‘ (4.63) t oes 4 phase! “FL 4.1.3.3 Two Identical Three-Phase Lines with Zero Sequence Coupling Only Just as @ transposed single-circuit three-phase line can usually be approximated as a balanced line, so two identical and parallel three-phase lines can often be approximated as “almost balanced” lines with an impedance matrix of the form *)To be consistent, lines with M= 1 and M= 2 are called “single-phase” and “two-phase” lines, respectively, in this manual. This differs from the IEEE Standards [76, p. 647], in which circuits with one phase conductor and one neutral conductor (which could be replaced by ground return), as well as circuits with two phase conductors and one neutral conductor (or ground return) are both called single-phase circuits for historical reasons. For Mp3, the definition in the IEEE Standards is the same as in this manual. zat a at eee a zat pp zon pp ou ee P 4-45 2 Pp zt ozs Da) eee pp (4.64) The transposition scheme of Fig. 4.22 would produce such a matrix form, which implies that the two circuits are only coupled in zero sequence, but not in positive or negative sequence. Such a complicated transposition scheme is seldom, if ever, used, but the writer suspects that positive and negative Sequence couplings in the more common double-circuit transposition scheme of Fig. 4.23 is often so weak that the model discussed here may be a useful approximation for the case of Fig. 4.23 as well. 7 Soa SE ei Fig. 4.22 - Double-circuit transposition scheme with zero sequence coupling only The matrix of Eq. (4.64) {8 diagonalized by modifying the transformation matrix of Eq. (4.58) to he le tou 1-4 o 0 o 0 oo (4.65) wu vee o -2 4-46 with [T]~! = [T]* again, which produces the diagonal matrix a (4.66) ay ay a If each circuit has three-phase sequence parameters Z' zero, Z' poss and if the three-phase zero sequence coupling between the “two circuits is Z',ero-couplingr then the ground mode G, inter-line mode IL and line mode L values required by the EMTP are found from Z'e = 2" zero + 2'zero-coupling’ 2"t, = Z' zero ~ 2'zero-coupling! (4.67a) 2, = 2"ponr with identical equations for the capacitances. The mode parameters can also be found from the matrix elements in Eq. (4.64), with Z'q = 2%, + 22", + 32"p, 2%, = By + 2D%_ - 32D, (4.67) Bi, = Dg - Bye If the two three-phase circuits are not identical, then the transformation matrix of Eq. (4.65) can no longer be used; instead, [T] depends on the particular tower configuration. 4.1.4 Symmetrical Components Symmetrical components are not used as such in the EMTP, except that the parameters of balanced lines after transformation to M-phase a, B, O-components are the same as the parameters of symmetrical components, namely zero and positive sequence values. The supporting routine LINE CONSTANTS does have output options for more detailed symmetrical component information, however, which may warrant some explanations. In addition to zero and positive sequence values, LINE CONSTANTS also prints full symmetrical component matrices. Its diagonal elements are the familiar zero and positive sequence values of the line; they are correct for the untransposed line, as well as for a line which has been balanced through proper transpositions. Off-diagonal elements are only meaningful for the 4-47 untransposed case, because they would become zero for the balanced line. For the untransposed case, these off-diagonal elements are used to define unbalance factors [47, p. 93]. The full symmetrical component matrices are no longer symmetric, unless the columns for positive and negative sequence are exchanged [27]. This exchange is made in the output of the supporting routine LINE CONSTANTS with rows listed in order “zero, pos, neg, and columns in order “zero, neg, pos, «..". With this trick, matrices can be printed in triangular form (elements in and below the diagonal), as is done with the matrices for individual and equivalent phase conductors. Syametrical components for two-phase lines are calculated with the transformation matrix of Eq. (4.63), while those of three-phase lines are calculated with a1 (phase! ~ (51 l%eyqq]* 84 [gyqq) = (S17 phase! (4.688) tdentical for currents zero here Ly, y where [veyanl * | Ypos Yneg i 1 1 1 1 1 L Is}=4 fir a2 a | ana ts t= 4]. a a2 (4.680) 1 a a2 1 az ia and a= 120 . * ‘The colums in these matrices are normalized”; in that form, [8] is unitary, ts} = (se) (4.69) where “*" indicates conjugate complex and “t" transposition, "ithe electric utility industry usually uses unnorealized transforeatton, fo which the factor for the [S]-matrix is 1 instead of 1/¥3, and for the [S]~*- matrix 1/3 instead of 1//3. The symmetrical component impedances are identical in both cases, but the sequence currents and voltages differ by a factor 73. 4-48 For M>3, the supporting routine LINE CONSTANTS assumes three-phase lines in parallel. Examples: M= 6: Two three-phase lines in parallel M= 9: Three three-phase lines in parallel M= 8: Two three-phase lines in parallel, with equivalent phase conductors no. 7 and 8 ignored in the transformation to symmetrical components. The matrices are then transformed to three-phase symmetrical components and mot_to M-phase symmetrical components of Eq. (4.62). For example for M= 6 (double-circuit three-phase line), iso [Is] 0 (23 yam) ; phase (4.70) 0 (s] 0 Is, with [S] defined by Eq. (4.68), Eq. (4.70) produces the three-phase symmetrical component values required in Eq. (4.67). Balancing of double-cireuit three-phase lines through transpositions never completely diagonalizes the respective symmetrical component matrices. The best that can be achieved is with the seldon-used transposition scheme of Fig. 4.22, which leads to Zero-t 0 9 Zzero-coupling ° u a Bost 0 ° o ° 0 0 Boer ° 0 ° (25 yma) . 2 ero-coupling 0 2 ‘Zero-II 0 0 o o 0 ° Zos-Il 0 ° 0 0 ° 2 os-IL (4.71) 4-49 If both circuits are identical, then z' =z =2 zero-I ~ “zero-II = 21,3 in that case, the transformation matrix of Eq. (4.65) can be used for diagonalization, The more common transposition scheme of Fig. 4.23 produces positive and zero sequence coupling between the two circuits (a) barrels rolled in (>) barrels rolled in opposite direction same direction. Fig. 4,23 - Double-circuit transposition scheme as well, with the nonzero pattern of the matrix in Eq. (4.71) changing to where "X" indicates nonzero terms. Re-assigning the phases in Fig. 4.23(b) to CI, BI, AL, AIL, BIT, CII from top to bottom would change the matrix further to cross-couplings between positive sequence of one circuit and negative sequence of the other circuit, and vice versa, 4-50 4.1.5 From the discussions of Section 4.1.3 it should have become obvious that the solution of M-phase transmission line equations becomes simpler if the M coupled equations can be transformed to M decoupled equations. These decoupled equations can then be solved as if they were single-phase equations. For balanced lines, this transformation is achieved with Eq. (4.58). Many lines are untransposed, however, or each section of a transposition barrel may no longer be short compared with the wave length of the highest frequencies occuring in a particular study, in which case each section must be represented as an untransposed line. Fortunately, the matrices of untransposed lines can be diagonalized as well, with transformations to “modal” parameters derived from eigenvalue/eigenvector theory. The transformation matrices for untransposed lines are no longer known a priori, however, and must be calculated for each particular pair of parameter matrices (21.501 and [Wnagel* To explain the theory, let us start again from the two systems of equations (4.31) and (4.32), Ven, ~ (Phe y= tz at dx ‘phase T nace! (4.728) and STohase [ae T= nase! Vpnase (4,720) with [age] * JolC',,4] 1F shunt conductances are ignored, as ts customarily done. By differentiating the first equation with respect to x, 4-51 and replacing the current derivative with the second equation, a second-order differential equation for voltages only is obtained, — phase ] = (2! lex phase! !¥pnase!!Ypnase!* (4.73a) Similarly, a second-order differential equation for currents only can be obtained, _ (a= estteneee IUzceecl omega de (4.736) where the matrix products are now in reverse order from that in Eq. (4.73a), and therefore different. Only for balanced matrices, and for the lossless high-frequency approxinations dicussed in Section 4.1.5.2, would the matrix products in Eq. (4.73a) and (4.73b) be identical. With eigenvalue theory, it becomes possible to transform the two coupled equations (4.73) from phase quantities to “modal” quantities in such a way that the equations become decoupled, or in terms of matrix algebra, that the associated matrices becone diagonal, e.g., for the voltages, ay [33 1] = Cayo! (4.74) ax with [A] being a diagonal matrix. To get from Eq. (4.73a) to (4.74), the phase voltages mist be transformed to mode voltages, with (phase! * (7,1 Vangel (4.754) and -1 Waoae * (Ty as (4.756) Then Eq. (4.73a) becomes [22 1 = 1 pase! (nase! (Ty! Maode! (4.76a) ae vw) (ohase! (nase! !Ty1 Maode which, when compared with Eq. (4.74), shows us that 4-52 UM = [nT (2 yage! (nase! Ty (4.760) nase! !¥pnase! 18 the eigenvalue/eigenvector problem. The diagonal elements of [A] are the etgenvalues of the matrix product [2),,.6][¥Snage!> amd [T,] 19 the matrix of eigenvectors or modal matrix of that matrix product. There are many methods To find the matrix [T,] which diagonalizes (2), for finding eigenvalues and eigenvectors. The most reliable method for finding the eigenvalues seems to be the QR-transformation due to Francis [3], while the most efficient method for the eigenvector calculation seems to be the inverse iteration scheme due to Wilkinson [4,5]. In the supporting routines LINE CONSTANTS and CABLE CONSTANTS, the "EISPACK"-subroutines [67] are used, in which the etgenvalues and eigenvectors of a complex upper Hessenberg matrix are found by the modified LR-method due to Rutishauser. This method is a predecessor of the QR~method, and where applicable, as in the case of positive definite matrices, is more efficient than the QR-method [68]. To transform the original complex matrix to upper Hessenberg form, stabilized elementary similarity transformations are used. For a given eigenvalue %,, the corresponding eigenvector [t,,] (= k-th column of [T,]) is found by solving the system of linear equations Lys hase! ~ %elU] Hey] = 0» (4.77) (nase with (U] = unit or identity matrix. Eq. (4.77) shows that the eigenvectors are not uniquely defined in the sense that they can be multiplied with any ) nonzero (complex) constant and still remain proper eigenvectors’, in contrast to the eigenvalues which are always uniquely defined. Floating-point overflow may occur in eigenvalue/eigenvector subroutines if the matrix is not properly scaled. Unless the subroutine does the scaling “This is important if matrices [T,] obtained from different programs are compared. The ambiguity can be removed in a number of ways, e.g., by agreeing that the elements in the first row should always be 1.0, or by normalizing the columns to a Euclidean vector length of 1.0, that is, by requiring tyjty,* + typtyy* + «+. = 1.0, with t* = conjugate complex of t. In the latter case, there is still ambiguity in the sense that each column could be multiplied with a rotation constant e/% and still have vector length = 1.0. 4-53 automatically, (2%nagq](¥nage! Should be scaled before the subroutine call, by dividing each element by -(ue,y,), a8 suggested by Galloway, Shorrows and Wedepohl [39]. This division brings the matrix product close to unit matrix, because [2,25] l¥ pase! 18 a diagonal matrix with elenents -we,u, if resistances, internal reactances and Carson's correction terms are ignored in Eq. (4.7) and (4.8), as explained in Section 4.1.5.2. The eigenvalues from this scaled matrix must of course be multiplied with ~u%e,y, to obtain the eigenvalues of the original matrix. In [39] it is also suggested to subtract 1,0 from the diagonal elements after the division; the eigenvalues of this modified matrix would then be the p.u, deviations from the eigenvalues of the lossless high-frequency approximation of Section 4.1.5.2, and would be much more separated from each other than the unmodified eigenvalues which lie close together. Using subroutines based on [67] gave identical results with and without this subtraction of 1.0, however. In general, a different transformation must be used for the currents, [ypase! 7 [7] Tgoae!> (4.788) and Upoae! 7 (TI pase! (4.780) ‘mode ‘phase! because the matrix products in Eq. (4.73a) and (4,73b) have different eigenvectors, though their eigenvalues are identical. Therefore, Eq. (4.73b) is transformed to ar a Se = UN pogo!» (4.79) ax’ with the same diagonal matrix as in Eq. (4.74). While [T,] is different from (t,], both are fortunately related to each other [58], (ty) = tr , (4,80) where “t" indicates transposition. It is therefore sufficient to calculate only one of them, Modal analysis is a powerful tool for studying power line carrier problems [59-61] and radio noise interference [62,63]. Its use in the EMTP 4-54 is discussed in Section 4.1.5.3. It is interesting to note that the modes in single-circuit three-phase lines are almost identical with the a, p, 0 components of Section 4.1.3.1 [58]. Whether the matrix products in Eq. (4.73) can always be diagonalized was first questioned by Pelissier in 1969 [64]. Brandao Faria and Borges da Silva have shown in 1985 [65] that cases can indeed be constructed where the matrix product cannot be diagonalized. It is unlikely that such situations will often occur in practice, because extremely small changes in the parameters (e-g-, in the 8th significant digit) seem to be enough to make it diagonalizable again. Paul [66] has shown that diagonalization can be guaranteed under simplifying assumptions, e-g-, by neglecting conductor resistances. The physical meaning of modes can be deduced from the transformation matrices [T,] and (T,]. Assume, for example, that column 2 of (T,] has entries of (-0.6, 1-0, -0.4). From Bq. (4.784) we would then know that mode-2 current flows into phase B in one way, with 60% returning in phase A and 40% returning in phase C. 4.1.5.1 Line Equations in Modal Domain With the decoupled equations of (4.74) and (4-79) in modal quantities, each mode can be analyzed as if it were a single-phase line. Comparing the modal equation 2, a Naode-k 4 y ae Ne Ymode-k with the well-known equation of a single-phase line, 2, av 2 =v, ae with the propagation constant y defined in Eq. (1.15), shows that the modal propagation constant y, is the square root of the eigenvalue, Ynode-k ~ “ke * 38y = Me a = attenuation constant of mode k (e.g, 1m Np/km), fmode-k with ay Py The phase velocity of mode k is = phase constant of mode k (e.g, in rad/km). 4-55 phase velocity = z, (4-82a) ik and the wavelength is wave length = 2 (4.82) Re While the modal propagation constant is always uniquely defined, the modal series impedance and shunt admittance as well as the modal characteristic impedance are not, because of the ambiguity in the eigenvectors. Therefore, modal impedances and admittances only make sense if they are specified together with the eigenvectors used in their calculation. To find them, transform Eq. (4-72a) to modal quantities Saode > (Re) = ry easel (I, (4.83) The triple matrix product is Eq. (4.83) is diagonal, and the modal series ‘node! impedances are the diagonal elements of this matrix Loge! * Uy) gles Ly (4.-84a) or with Eq- (4-80), . =r [2yoae) 7 IT") (nage! (TL) (4.840) Similarly, Eq. (4.72b) can be transformed to modal quantities, and the modal shunt admittances are then the diagonal elements of the matrix cr ote! "ET" nage! (ys (6.858) or with Eq. (4.80), Treen a teeg)) (nese It ryle (4.850) phase The proof that both (Z1,4,] and [¥1,4,] are dfagonal is given by Wedepohl [58]. Finally, the modal characteristic impedance can be found from the scalar equation 1 aode-k Zohar-mode-k - (4.868) or from the simpler equation 4-56 Ynode-k ?char-mode-k ~ Toa (486d) A good way to obtain the modal parameters may be as follows: First, obtain the eigenvalues A, and the eigenvector matrix [T,] of the matrix Product (2) s5¢] [snagel* Then find [Y104,] from Eq. (4.85b), and the modal series impedance from the scalar equation Xe z =k. G86) mode-k ~ Yoo The modal characteristic impedance can then be calculated from Eq. (4.86a), or from Eq. (4.86b) if the propagation constant from Eq. (4.81) is needed as well. If [T,] is needed, too, it can be found efficiently from Eq. (4.85a) (T 1) = [pagel (Ty) Mhoael > (4-85¢) because the product of the first two matrices is available anyhow when ] is found, and the post-multiplication with (¥i,4 J"! is simply a Mode! ode! multiplication of each coluan with a constant (suggested by Luis Marti). Eq. (4.85¢) also establishes the link to an alternative formula for [T,) mentioned in [57], (Wrpase! [Ty 1D, with [D] being an arbitrary diagonal matrix. Setting (D] = (¥t,4,]7) leads us to the desirable condition [T,] = cas ie of Eq. (4.80). If the unit matrix were used for (D], all modal matrices in Eq. (4.84) and (4.85) would still be diagonal, but with the strange- looking result that all modal shunt adoittances becone 1.0 and that the modal series impedances become \,- Eq+ (4.80) would, of course, no longer be fulfilled. For a lossless line, the nodal series impedance would then becone a negative resistance, and the modal shunt admittance would become a shunt conductance with a value of 1.0 S. As long as the case is solved in the frequency donain, the answers would still be correct, but 1t would obviously be wrong to associate such modal parameters with - Beet, na- Bec (with R’ negative and G' = 1.0) in the time domain. 4-57 4.1.5.2 Lossless High-Frequency Approximation In lightning surge studies, many simplifying assumptions are made, For example, the waveshape and amplitude of the current source representing the Lightning stroke is obviously not well known. Similarly, flashover criteria in the form of volt-time characteristics or integral formulas [8] are only approximate. In view of all these uncertainties, the use of highly sophisticated line models is not always justified. Experts in the field of lightning surge studies normally use a simple line model in which all wave speeds are equal to the speed of light, with a surge impedance matrix [2surge! im Phase quantities, where 244 surge 7 60 8nC2hy/r,), (4,878) Zsx-surge ~ 69 ND, /d,)» (4,876) all wave speeds = speed of light, (4,876) with r, being the radius of the conductor, or the radius of the equivalent conductor from Eq. (4.30) in case of a bundle conductor”). Typically, each span between towers is represented separately as a line, and only a few spans are normally modelled (3 for shielded lines, or 18 for unshielded lines in (8]). For such short distances, losses in series resistances and differences in modal travel tines are negligible. ‘The effect of corona is sometimes included, however, by modifying the simple model of Bq. (4.87) [8]. It is possible to develop a special line model based on Eq. (4.87) for the EMTP, in which all calculations are done in phase quantities. But as shown here, the simple model of Eq. (4.87) can also be solved with modal parameters as a special case of the untransposed line, The simple model follows from Eq. (4.72) by making two assumptions for a “lossless high- frequency approximation" : 1, Conductor resistances and ground return resistances are ignored. 2, ‘The frequencies contained in the lightning surges are so high that all currents flow on the surface of the conductors, and on the surface of the earth. ‘Ground wires are usually retained as phase conductors in such studies. If they are eliminated, the method of Section 4.1.2.1 must be used on [Zoyrgq] ‘surge: 4-58 Then the elenents of (2),,,,4] become 1. Bo . Yo 2 gt So ge tm (Angle), 2 gy = Sw ye An Dy /dyy)s (4.88) while [Y'] 1s obtained by inverting the potential coefficient matrix, Gey so et, (4.89) with the elements of [P’] being the same as in Eq. (4.88) if the factor Jong/(2x) ts replaced by 1/(2ne,)- Then both matrix products in Eq. (4.73) become diagonal matrices with all elenents being dete wet KE Ly oe Me (4.90) otto These values are automatically obtained from the supporting routines LINE CONSTANTS and CABLE CONSTANTS as the eigenvalues of the matrix products tn Eq. (4.73), by simply using the above two assumptions in the input data (all conductor resistances = 0, GMR/r = 1.0, no Carson correction terms). The calculation of the eigenvector matrix [Ty] or [T,] needed for the untransposed line model of Section 4.2 breaks down, however, because the matrix products in Eq. (4.73) are already diagonal. To obtain [T,], let us first assume equal, but nonzero conductor resistances &'. Then the eigenvectors (t,,] are defined by 2, eppeyt = Pell + Suk EP OTe) Ale yd (491) with the expression in parenthesis being the matrix product [2 tv 1, phase’! "phase! and (U] = unit matrix. Eq. (4.91) can be rewritten as te) fed = Me-modi tied! "vil > (4.92) with odifted etgenveluce . wed HB he moastted "Me * Foto (93) Eq. (4.92) is valid for any value of R', including zero. It therefore follows that [T,] is obtained as the eigenvectors of [?']"!, or alternatively as the eigenvectors of [P'] since the efgenvectors of a matrix are equal to 4-59 the eigenvectors of its inverse. The eigenvalues of [P"]"! are not needed because they are already known from Eq. (4.90), but they could also be obtained from Bq. (4.93) by setting R' = 0. For this simple model, [T,] 1s @ real, orthogonal matrix, (rylti* = 0) (4.94) and therefore, i) = (ty) 4.95) D.E. Hedman has solved this case of the lossless high-frequency approximation more than 15 years ago [45]. He recommended that the eigenvectors be calculated from the surge impedance matrix of Eq. (4.87), which 1s the sane as calculating them from [P'] since both matrices differ only by a constant factor. one can either modify the line constants supporting routines to find the eigenvectors from (P'] for the lossless high-frequency approximation, as was done in UBC's version, or use the same trick employed in Eq. (4.91) in an unmodified program: Set all conductor resistances equal to some nonzero value R', set GMR/r=1, and ask for zero Carson correction terms. If the eigenvectors are found from {P'], then it is advisable to scale this matrix first by mltiplying all elements with 2ne,. The lossless high-frequency approximation produces eigenvectors which differ from those of the lossy case at very high frequencies [61]. This ts unimportant for lightning surge studies, but important for power line carrier problens. Example: For a distribution line with one ground wire (Fig. 4.24) the lossless high-frequency approximation produces the following modal surge impedances and transformation matrix, mode | Zsurge-mode (2) L 993.44 2 290.67 5) 360.70 4 310.62 4-60 Fig. 4.24 - Position of phase conductors A,B,C and ground wire D (average height, all dimensions in a). Conductor diameter = 10.1092 mm 0.52996 0.82860 -0.18049 0 (T,] = (ty) = ]0-49080 -0.21322 0.46222 -0.707 11 0.49080 0.21322 0.46222 0.70711, 0.48721 -0.47170 -0.73493, ° Converted to ph (ty2 t ‘surge-modellT,] 9 OF 490.33 ]= [176.95 484.89 176.95 174.27 190.74 144.26 (2, ‘eurge-phase’ e quantities, the surge impedance matrix becomes symmetric Q 484.89 144,26 495.31 The elements from Eq. (4.87) are slightly larger, by a factor of 300 000/299 792, because the supporting routine LINE CONSTANTS uses 299 792 km/s for the speed of light, versus 300 000 km/s implied in Eq- (4.87). 4-61 ‘The tranemiesion line part of an EMTP input file for modelling 4 spans, each 90 m long, would look as follows: -11A 2A 993.A40.3E-S 204 218 (2B 290.470. 3E-6 2 sic 2c 360.700. 3E-6 2 ~41D 2D 310.620. 3E-6 2 0.529956 0.828595 0. 180491 o. 3 0.490801 0.213224 0462222 -0.707107 0.490801 0.213224 0.462222 04707107 Z 0.487211 0.471701 -0.734931 O - -128° 3A ia 2A -228 (3B % wee 42D 03D : “13a 4A 1A 4 2 -23B 4B upc VERSION INPUT; MAY DIFFER |“ “3300 4c SLIGHTLY PROM BPA veRSion npur | $s -43D 04D g Sian | Sal in| 20 & -24B SB + -34c SC -440 SD 7 11D 106 390.0 .03E-6 2) gz “120 2DG 1D 106 . gs wed “13D «-3DG1D«1G 2 i “14D 406, 1D. : 24 et “15D SDG «1D «1G un Sp ae 1D6 20. 206 16 Lower Rooting resistomees, 406 106 QOR WAN; NBR ind). sD 1D6 : 36 168 The modelling of long lines as coupled shunt resistances [R] = [Zyurge-phase! BA# already been discussed in Section 3.1.3. In the example above, such a shunt resistance matrix could be used to represent the rest of the line after the 4 spans from the substation. Simply using the 4 x 4 matrix would be unrealistic with respect to the ground wire, however, because 4t would imply that the ground wire is ungrounded on the rest of the line. More realistic, though not totally accurate, would be a 3 x 3 matrix obtained so! 20d [¥tnggq] in which the ground wire has been eliminated. This model implies zero potential everywhere on the ground wire, in contrast from [2a 462 to the four spans where the potential will more or less oscillate around zero because of reflections up and down the towers. For EMIP users who are reluctant to Comparison with More Accurate Modell use the simple model described in this section, a few comments are in order. First, let us compare exact values with the approximate values. If we use constant parameters and choose 400 kHz as a reasonable frequency for Lightning surge studies, then we obtain the results of table 4.5 for the test example above, assuming T/D = 0.333 for skin effect correction and internal inductance calculation with the tubular conductor formula, R' 4, ~ 0.53609 Q/km, and p= 100 sm. Table 4.5 - Exact line parameters at 400 kiz mode | Zyurge-aode (9) | wave velocity (m/s) | RC s/km) 1 1027.6-$33.9 285.35 597.4 2 292.0-40.5 299.32 7.9 3 361.9-J0.5 299.37 8.2 4 311.1-J0.5 299.32 8.0 ‘The differences are less than 0.5% in surge impedance and wave speed for the aerial modes 2 to 4, and not more than 5% for the ground return mode 1. ‘These are small differences, considering all the other approximations which are made in lightning surge studies. If series resistances are included by lumping then in 3 places, totally erroneous results may be obtained if the user forgets to check whether R/4 < 2, in the ground return mode. For the very short Line Length of 90 in this example, thts condition would still be fulfilled here. Using constant paraneters at a particular frequency 19 of course an spproxiaation as well, and sone users may therefore prefer frequency dependent sodels. For very short Line lengths, auch as $0 m in the example, ost frequency-dependent models are probably unreliable, however. It say therefore be more sensible to use the simple model described here, for which 4-63 answers are reliable, rather than sophisticated models with possibly unreliable answers. ‘A somewhat better lossless line model for lightning surge studies than the preceding one has been suggested by V. Larsen [92]. To obtain this better model, the line paraneters are first calculated in the usual way, at a certain frequency which 1s typical for lightning surges (e.g-, at 400 kHz). The resistances are then set to zero when the matrix product [2" ase] %' jpase! 18 formed, before the modal parameters are computed. With this approach, [7,] will always be real. Table 4.6 shows the nodal parameters of this better lossless model. They differ very little from those in Table 4.5. Table 4.6 - Approximate modal parameters at 400 kHz with ReO mode (a) | wave velocity (m/s) z, ‘surge-mode In particular, the wave velocity of the ground return mode 1 is now much closer to the exact value of Table 4.5. The transformation matrix which goes with the modal parameters of Table 4.6 is 0.40795 0.84115 -0.22316 o Im] = | 0.55628 -0,18448 0.44910 -0.70711 0.55628 0.18448 0.44910 0.70711 0.46335 -0.47371 -0.73947 ° In this case [T,] 16 no longer equal to [T,]; Eq. (4.80) must be used instead. 4-64 4.1.5.3 Approximate Transformation Matrices for Transient Solutions ‘The transformation matrices (T,] and [T,] are theoretically complex, and frequency-dependent as well. With a frequency-dependent transformation matrix, modes are only defined at the frequency at which the transformation matrix was calculated. Then the concept of converting a polyphase line into decoupled single-phase lines (in the modal domain) cannot be used over the entire frequency range. Since the solution methods for transients are much simpler if the modal composition is the same for all frequencies, or in other words, if the transformation matrices are constant with real coefficients, it 4 worthwhile to check whether such approximate transformation matrices can be used without producing too much error. Fortunately, this is indeed possible for overhead lines [66,78]. Guidelines for choosing approximate (real and constant) transformation matrices have not yet been worked out at the time when these notes are being written. The frequency-dependent line model of J. Marti discussed in Section 4.2.2.6 needs such a real and constant transformation matrix, and wrong answers would be obtained if a complex transformation matrix were used instead. Since a real and constant transformation matrix is always an approximation, its use will always produce errors, even if they are small and acceptable. The errors may be omall in one particular frequency region, and larger in other regions, depending on how the approximation is chosen. One choice for an approximate transformation matrix would be the one used in the lossless approximations discussed in Section 4.1.5.2. This may be the best choice for lightning surge studies For switching surge etudies and similar types of studies, the preferred approach at thie time seems to be to calculate (T,] at a particular frequency (e.g, at 1 kz), and then to ignore the imaginary part of it. In this approach, [T,] should be predominantly real before the imaginary part is discarded. One cannot rely on this when the subroutine returns the 450° constant would still be a proper eigenvector. Therefore, the columns of [T,] should be normalized in such a way that its components lie close to the real eigenvectors, since an eigenvector multiplied with e”” or any other 4-65 axis. one such normalization procedure was discussed by v. Brandwajn [79]. The writer prefers a different approach, which works as follows: aanore shunt conductances, as is customarily done. ‘Then [Y'yhase] is purely imaginary. Use Eq.(4.85) to find the diagonal elements of the modal shunt admittance matrix Y'mode preliminary* 2. In general, these "preliminary" modal shunt admittances will not be purely imaginary, but J0C'noge-,'exP(-Jja,) instead. Then normalize [T,] by multiplying each column’ with exp(ja;/2). With this normalized transformation matrix, the modal shunt admittances will become joC'poge-,, OF purely imaginary as in the phase domain. 3. To obtain the approximate (real and constant) transformation matrix, discard the imaginary part of the normalized matrix from step 2. Use the approximate matrix [Ty-approx.] from step 3 to find modal series impedances and modal shunt admittances from Eq. (4.84) and (4.85) over the frequency range of interest. If [T,] is needed, use crt, (4.96) (Ty-approx.] = [T*y-approx.] If the line model requires nonzero shunt conductances, add them as modal parameters. Usually, only conductances from phase to ground are used (with phase-to-phase values being zero); in that case, the modal conductances are the same as the phase-to- ground conductances if the latter are equal for all phases. In the UBC Line Parameters Program, and in Version 2.0 of the DCG/EPRI EMTP, the normalization is done on [T,] rather than on [T,], by using [P'pnasel=[C'phase) 1 and making sure that [P'poge] becomes purely real. in stép 3, [Tj approg.] would be obtained “by discarding th Imaginary pare #fon (T(}"SF step 2, and in step,4, [Tyeapprox.] Wowld be calculated from (Ty approx.] = (T*i-approx.] *+ An interesting method for finding approximate (real and constant) transformation matrices has been suggested by Paul [66]. By ignoring conductor resistances, and by assuming that the Carson correction terms AR", + JAX", in Eq. (4.7) and AR'y, +)AX"y, in Eq. (4.8) are all equal (all élements in the matrix of correction terms have one and the same value), the approximate transformation matrix [Ty-approx.] is obtained as the eigenvectors of the matrix product 4-66 U1... (Ci nase! : Ph Le. with all elements of the second matrix being 1. To find [T, appro, ]> Eq. (4.96) would have to be used. Wasley and Selvavinayagamoorthy [93] find the approximate transformation matrices by simply taking the magnitudes of the complex elements, with an appropriate sign reflecting the values of their arguments. They compared results using these approximate matrices with the exact results (using complex, frequency-dependent matrices), and report that fairly high accuracy can be obtained if the approximate matrix is computed at a low frequency, even for the case of double-circuit lines. If the M-phase line is assumed to be balanced (Section 4.1.3.2), then the transformation matrix is always real and constant, and known a priori with Eq. (4.58) and Eq. (4-59). Two identical and balanced three-phase lines with zero sequence coupling only have the real and constant transformation matrix of Eq. (4.65). 4.2 Line Models in the EMTP The preceding Section 4.1 concentrated on the line parameters per unit length. These are now used to develop line models for lines of a specific length. For steady-state solutions, lines can be modelled with reasonable accuracy as nominal mcircuits, or rigorously as equivalent tcircuits. For transient solutions, the methods become more complicated, as one proceeds from the simple case of a single-phase lossless line with constant parameters to the more realistic case of a lossy polyphase line with frequency-dependent parameters. 4-67 Lines can be represented rigorously in the steady-state solution with exact equivalent x-circuits. Less accurate representations are sometimes used, however, to match the model to the one used in the transient simulation (e-g-, lumping R in three places, rather than distributing it evenly along the line, or using approximate real transformation matrices instead of exact complex matrices). For lines of moderate lectrical” length (typically < 100 kn at 60 Hz), nominal x-circuits are often accurate enough, and are probably the best models to use for steady-state solutions at power frequency. If the steady-state solution 1s followed by a transient simulation, or if steady-state solutions are requested over a wide frequency range, then the nominal m-circuit mist either be replaced by a cascade connection of shorter nominal n-circuits, or by an exact equivalent n-circuit derived from the distributed parameters. 4.2.1.1. Nominal M-Phase x-Circuit For the nominal M-phase n-circuit of Fig. 3-10, the series impedance matrix and the two equal shunt susceptance matrices are obtained from the per unit length matrices by simply multiplying them with the line length, as shown in Eq. (4-35) and (4.36). ‘The equations for the coupled lumped elements of this M-phase n-circuit have already been discussed at length in Section 3, and shall not be repeated here. Nominal n-circuits are fairly accurate if the line is electrically short. This is practically always the case if complicated transposition schemes are studied at power frequency (60 Hz or 50 Hz). Fig. 4.25 shows a typical example, with three circuits on the same right-of-way. In this case, 5-6) would be each of the five transposition sections (1-2, 2-3, 3-4, 4~ represented as a nominal 9-phase n-circutt. While a nominal n-circuit would already be reasonably accurate for the total line length of 95 ka, nominal n- circuits are certainly accurate for each transposition section, since the longest section is only 35 km long. A comparison between measurements on the 4-68 distances in km |}-——— 94.9 1.6 a.s| | 20.9] 32.2 | 35.4 oe cme co 550 kV under construction 550 kv in operation y 360 kv in operation phase Fig. 4.25 - Transposition scheme for three adjacent circuits Table 4.7 - Comparison between measurements and EMIP results (voltages on energized line Ll = 372 kV and on L2 = 535 kV) phase | measurement | EMTP results Induced voltages on A 30 kv 27.5 kV de-energized line 13 if open B 15 kv 13.8 kV at both ends c 10 kv 7.8 kV Grounding currents if A MLA 10.5 A de-energized line 13 is B 5A 3.24 grounded at right end c TA 1.5 A de-energized line L3 and computer results is shown in table 4.7 [80]. The coupling in this case is predominantly capacitive. Because nominal x-circuits are so useful for studying complicated transposition schemes, a “CASCADED PI"-option was added to the BPA EMIP. 4-69 With this option, the entire cascade connection is converted to one single x- cireuit, which is an exact equivalent for the cascade connection. This is done by adding one “component” at a time, as shown in Fig. 4.26. The aN 3) Q Bquivalent Equivalent for = for components components Lp 2eveeeeKeL, 1,2,. x Fig. 4.26 - Schematic illustration of cascading operation for K-th component “component” may either be an M-phase x-circuit, or other types of network elements such as shunt reactors or series capacitors. Whenever component k is added, the nodal admittance matrix for nodes 1, 2, 3 is reduced by eliminating the inner nodes 2, to form the new admittance matrix of the equivalent for the cascaded components 1, 2, ... K. This option keeps the computational effort in the steady-state solution as low as possible by not having to use nodal equations for the inner nodes of the cascade connection, at the expense of extra computational effort for the cascading procedure. e Lines 4.2.1.2 Equivalent x-Cireuit for Single-Phi Lines defined with distributed parameters at input time are always converted to equivalent x-circuits for the steady-state solution. For lines with frequency-dependent parameters, the exact equivalent 1- circuit discussed in Section 1 1s used, with Bq. (1-14) and (1.15). The same exact equivalent n-circuit is used for distortionless and lossless line models with constant parameter In many applications, line models with constant parameters are accurate enough. For example, positive sequence resistances and inductances are 4-70 fairly constant up to approximately 1 kz, as shown in Fig. 4.20. But even with constant parameters, the solution for transients becomes very complicated (except for the unrealistic assumption of distortionless propagation). Fortunately, experience showed that reasonable accuracy can be obtained if L' and C' are distributed and if R= RE (4.97) 4s lumped in a few places as long as R << 7,.,,,+ In the EMIP, R/? 1s lumped in the middle and R/4 at both ends of an otherwise lossless line, as shown in Fig. 4.27, and as further discussed in Section 4.2.2.5. For this aps po ais 1 2 3 4 lossless lossless line line Fig. 4.27 - Line representation with lumped resistances transient representation, the EMTP uses the sane assumption” in the steady- state solution, to avoid any discrepancies between ac steady-state initialization and subsequent transient simulation, even though experiments have shown that the differences are extremely small at power frequency. By using equivalent x-circuits for each lossless, half-length section in Fig. 4.27, and by eliminating the “inner” nodes 1, 2, 3, 4, an equivalent x- circuit (Fig. 1.2) was obtained by R.M. Hasibar with 2 = R cos’ur - (0.5 + 0.03125 35) R sin’ur + J simur cosur + z 2 (0.375 | + 22), series (4.98) *) The EMTP should probably be changed to by-pass this option if only steady state solutions are requested, either at one frequency or over a range of frequencies. a7 sinut + # sin ar cos ut hy 2 shunt series where 1 = length 7", 2-/E, (4.99) R= length + RB" 4.2.1.3 Equivalent M-Phase 1-Circuit To obtain an equivalent M-phase mcircuit, the phase quantities are first transformed to modal quantities with Eq. (4.84) and (4.85) for untransposed lines, or with Eq. (4.58) and (4.59) for balanced lines. For identical balanced three-phase lines with zero sequence coupling only, Eq- (4.65) 1s used. For each mode, an equivalent single-phase m-circuit is then found in the same way as for single-phase lines; that is, either as an exact equivalent wcircuit with Eq. (1-14) and (1.15), or with Eq. (4.98) and (4.99) for the case of lumping R in three places. These single-phase modal reircuits each have a series admittance Y.o24.5 node and two equal shunt shunt-mode’ BY assembling these admittances as diagonal matrices, the admittance matrices of the M-phase wcircuit in phase adeiveances } quantities are obtained fron ied J (tye on (4.100) series series-mode and 1 © F Mohune! = (74) shune-model [4] (4.101) While it is always possible to obtain the exact equivalent M-phase m™ circutt at any frequency in this way, approximations are sometimes used to 4-72 match the representation for the steady-state solution to the one used in the transient solution. One such approximation is the lumping of resistances as shown in Fig. 4.27. Another approximation is the use of real and constant fed in Section transformation matrices in Eq. (4-100) and (4.101), as discu 4.1.5.3. 4.2.2 Transient Solutions Historically, the first 1ine models in the EMTP were cascade connections of meireuits, partly to prove that computers could match switching surge study results obtained on transient network analyzers (TNA's) at that time. On TNA's, balanced three-phase lines are usually represented with decoupled ‘conductor meireults, as shown in Fig. 4.28. This representation can aL ‘pos, 22 2, BL ‘BOS. B2 ML Nz Fig. 4.28 - Four-conductor mcircuit used on TNA's easily be derived from Eq. (4.44) by rewriting it as -av, <@ +2 Gy +i, +1) (4-102) & for phase A, and similar for phases B and C. The first term in Eq. (4-102) ts 2" ot, (or branch Al-A2 in Fig. 4.28), while the second tera Le the 4-73 common voltage drop caused by the earth and ground wire return current I, + I, + Ig (branch NI-N2 in Fig. 4,28). Note, however, that Fig. 4.28 is only valid if the sum of the currents flowing out through a line returns through the earth and ground wires of that same line. For that reason, the neutral nodes N2, N3, «+. must be kept floating, and only NI at the sending end is grounded. Voltages with respect to ground at location i are obtained by measuring between the phase and node N,. In meshed networks with different R/X-ratios, this assumption is probably not true. For this reason, and to be able to handle balanced as well as untransposed lines with any number of phases, M-phase rcircuits were modelled directly with M x M matrices, as discussed in Section 4.1.2.4. Voltages to ground are then simply the node voltages. Comparisons between these M-phase wcircuits, and with the four- conductor meircuits of Fig. 4.28 confirmed that the results are identical. ‘The need for travelling wave solutions first arose in connection with rather simple lightning arrester studies, where lossless single-phase line models seemed to be adequate. Section 1 briefly discusses the solution method used in the EMP for such lines. This method was already known in the 1920's and 1930's and strongly advocated by Bergeron [81]; it is therefore often called Bergeron's method. In the mathematical literature, it is known as the method of characteristics, supposedly first described by Riemann. Te soon became apparent that travelling wave solutions were mich faster and better suited for computers than cascaded circuits. To make the travelling wave solutions useful for switching surge studies, two changes were needed from the simple single-phase lossless line: First, losses had to be included, which could be done with reasonable accuracy by simply lumping R in three places. Secondly, the method had to be extended to M-phase lines, which was achieved by transforming phase quantities to modal quantities. originally, this was limited to balanced lines with built-in transformation matrices, then extended to double-circuit lines, and finally generalized to untransposed lines. Fig. 4.29 compared EMTP results with results obtained on TNA's, using the built-in transformation matrix for balanced three-phase Lines and simply lumping R in three places. 4-14 e 7__length=202.8 km X =X 5127 2 2,70.04+50.318 9/km 2,70.26+51.015 8/em €=11.86 nF/km —C,=7.66 nF/km | [| 20 40 —— t (ms) Fig. 4.29 - Energization of a three phase line. Computer simulation results (dotted line) superimposed on 8 transient network analyzer results for receiving end voltage in phase B. Breaker contacts close at 3.05 ms in phase A, 8.05 ms in phase B, and 5.55 ms in phase C (t=0 when source voltage of phase A goes through zero from negative to positive) [82]. Reprinted by permission of CIGRE While travelling wave solutions with constant distributed L', C’ and constant lumped R produced reasonably accurate answers in many cases, as shown in Fig. 4.29, there were also cases where the frequency dependence, especially of the zero sequence impedance, could not be ignored. Choosing constant line parameters at the dominant resonance frequency sometimes Amproved the results. Eventually, frequency-dependent line models were developed by Budner [83], by Meyer and Dommel [84] based on work of Snelson [85], by Semlyen [86], and by Ametani [87]. A careful re-evaluation of frequency-dependence by J. Marti [88] led to a fairly reliable solution method, which seems to become the preferred option as these notes are being written. J. Marti's method will therefore be discussed in more detail. 475 4.2.2.1 Nominal Circuits Nominal n-circuits are generally not the best choice for transient solutions, because travelling wave solutions are faster and usually more accurate. Cascade connections of nominal n-circuits may be useful for untransposed lines, however, because one does not have to make the approximations for the transformation matrix discussed in Section 4.1.5.3. On the other hand, one cannot represent frequency-dependent line parameters and one has to accept the spurious oscillations caused by the lumpiness. Fig. 4.30 shows these oscillations for the simple case of a single-phase 20 vw distributes paraneters 1 ge tectsoulse 3 toate fs ° 8 a0 Fig. 4.30 - Voltage at receiving end of a single phase line if a de voltage of 10 V is connected to the sending end at t=0 (line data: R = 0.0376 G/km, L = 1,52 mi/km, C= 14.3 nF/km, length 320 km; receiving end terminated with shunt inductance of 100 mii; line data could represent the zero-sequence mode of a double- circuit line) 4-76 line being represented with 8 and 32 cascaded nominal circuits. The exact solution with distributed parameters is shown for comparison purposes as well. The proper choice of the number of mcircuits for one line is discussed in [89], as well as techniques for damping the spurious oscillations with damping resistances in parallel with the series R-L branches of the mcircuit of Fig. 4.28. The solution methods for nominal mcircuits have already been discussed in Section 3.4, With M-phase nominal recircuits, untransposed lines (or sections of a line) are as simple to represent as balanced lines. In the former case, one simply uses the matrices of the untransposed line, whereas in the latter case one would use matrices with averaged equal diagonal and averaged equal off-diagonal elements. 4.2.2.2 Single-Phase Lossless Line with Constant L' and C’ The solution method for the single-phase case has already been explained in Section 1. The storage scheme for the history terms is the same as the one discussed in the next Section 4.2.2.3 for M-phase lossless lines, except that each single-phase line occupies only one section in the table, rather than M sections for M modes. Similarly, the initialization of the history terms for cases starting from linear ac steady-state initial conditions is the same as in Eq. (4.108). The solution is exact as long as the travel time 1 is an integer multiple of the step size At. If this is not the case, then linear interpolation is used in the ENTP, as indicated in Fig. 4.31. Linear interpolation is believed to be a reasonable approximation for most cases, since the curves are usually smooth rather than discontinuous. If discontinuities or very sharp peaks do exist, then rounding 1 to the nearest integer multiple of At may be more sensible than interpolation, however. There is no option for this rounding procedure in the EMTP, but the user can easily accomplish this through changes in the input data. Fig. 4.32 compares results for the case of Fig. 4.30 with sharp peaks with and without linear 4-77 hist vw) -18.1— 1 Fig. 4.32 - Effect of linear interpolation on sharp peaks. Dotted line: At = 9-324...ps to make + integer multiple of At. Solid line: At = 10 us (+ not integer mitiple of at) interpolation. The line was actually not lossless in this case, but the losses were represented in a simple way by subdividing the line into 64 4-78 Lossless sections and lumping resistances in between and at both ends. The interpolation errors are more severe if lines are split up into many sections, as was done here. If the line were only split up into two lossless sections, with R lumped in between and at both ends, then the errors in the peaks would be less (the First peak would be 18.8, and the second peak would be -15.4). The accumulation of interpolation errors on a line broken up into many sections, with + of each section not being an integer multiple of At, can easily be explained. Assume that a triangular pulse is switched onto a long, lossless line, which 1s long enough so that no reflections come back from the remote end during the time span of the study (Fig. 4.33). Let us ie! A vit) f Foe + oe Fig. 4.33 - Single-phase lossless line energized with triangular pulse look at how this pulse becomes distorted through interpolation as it travels down the line tf (a) the line is broken up into short sections of travel time 1.5 At each, and (b) the Line from the sending end to the measuring point is represented as one section (r= k + 1.5 At, with k= 1, 2, 3,...)- At any point on the line, the current will be 197%, and between points 1 and 2 separated by + (Fig. 4.33), va(t + ) = vile). 4-79 This last equation was used in Fig. 4.34, together with linear interpolation, to find the shape of the pulse as it travels down the line. The pulse loses its amplitude and becomes wider and wider if it is broken up into sections of travel time 1.5 At each. On the other hand, the pulse shape never becomes as badly distorted if the line is represented as one single section. What are the practical consequences of this interpolation error? Table 4.8 compares peak overvoleages fron a BEA svitching surge study on « 1200 KV three-phase line, 133 miles long. Each section was split up into Table 4.8 - Interpolation errors in switching surge study with At = 50 us Peak overvoltages (MV) Run | Line model 7 phase A | phase B | phase C 1 |single section} 1-311 1.191 1.496 2 7 sections 1.276 1.136 1.457 3 |single section] 1.342 1.167 1.489 with t rounded two lossless half-sections, with R lumped in the middle and at both ends, as explained in Section 4.2.2.4. Run no. 1 shows the results of the normal line representation as one section. Run no. 2 with subdivision into 7 sections produces differences of 2.6 to 4.7%. In run no. 3 the zero and positive sequence travel times t, = 664.93 us and + a and 450 us, respectively, to make them integer multiples of At = 50 us. = 445.74 ys were rounded to 650 ‘These changes could be interpreted as a decrease in both L', and C', of 2.25%, and as an increase in both L', and C', of 0.96%, with the surge *) The problem of interpolation errors is basically the same for single phase and M-phase lines; therefore, a three-phase case is presented here for which data was already available. Choosing a step size At which makes the travel time t an integer multiple of At is more difficult for three-phase lines, however, because there are two travel times for the positive and zero sequence mode on balanced lines (or three travel times for the 3 modes on untransposed lines). 4-80 o rs) LINE FROM SENDING END 70 x LINE FROM SENDING EXD 70 x BROKEN UP INTO SECTIONS OF REPRESENTED AS ONE. SINGLE res be SECTION ae sending end 1.5 8 down the Line La La wa" LA A A 4.5 at down the Line 6 Bt down the Line 7.5 b¢ down the Line 9 fe down the Line Fig. 4.34 - Pulse at incremental distances down the line 4-81 impedances remaining unchanged. Since Mine paraneters are probably no more accurate than 45% at best anyhow, these implied changes are quite acceptable. With rounding, a slightly modified case 1s then solved without interpolation errors. Whether an option for rounding 1 to the nearest integer multiple of ‘At should be added to the EMIP is debatable. In general, rounding may imply much larger changes in L', C' than in this case, and if implemented, warning messages with the magnitude of these implied changes should be added as well. In Table 4.8, runs no. 3 and 1 differ by no more than 2.3%, and the interpolation error is therefore acceptable if the line is represented as one section. Breaking the line up into very many sections may produce unacceptable interpolation errors, however. If the user is interested in a “voltage profile” along the line, then a better alternative to subdivisions into sections would be a post-processor “profile program” which would calculate voltages and currents at intermediate points along the line from the results at both ends of the line. Such a program is easy to write for lossless and distortionless lines. Luis Marti developed such a profile algorithm for the more complicated frequency~ dependent line models, which he merged into the time step loop of the ENTP [90]. This was used to produce movies of travelling waves by displaying the voltage profile at numerous points along the line at time intervals of At. Fig. 4.34(a) suggests a digital filtering effect from the interpolation which is similar to that of the trapezoidal rule described in Section 2.2.1. To explain this effect, Eq. (1.6) must first be transformed from the time domain Zvjcey - g(t) =F vg td + ty (t-0 into the frequency domain, 1 “Jur r-7\ lat Im) ve (4.103) ink) For simplicity, let us assume that voltage and current phasors V, and T,, at y, ke node m are known, and that we want to find 1=7~~ I, at node k. Without 4-82 interpolation errors, Eq. (4-103) provides the answer. If interpolation is used, and if for the sake of simplicity we assume that the interpolated value lies in the middle of an interval At, then Eq. (4.103) becomes ye (#3) gw (r= FH) +e “tle ) C4104) Tracerpotated Therefore, the ratio of the interpolated to the exact value becomes 1, jnkerpolated ~ cos (y AE), (4.105) ‘exact which is indeed somewhat similar to Fig. 2.10 for the error produced by the trapezoidal rule. Single-phase lossless line models can obviously only approximate the complicated phenomena on real lines. Nonetheless, they are useful in a number of applications, for example (a) in simple studies where one wants to gain insight into the basic phenomena, (>) in lightning surge studies, and (c) as a basis for more sophisticated models discussed later. For lightning surge studies, single-phase lossless line models have been used for a long time. They are probably accurate enough in many cases because of the following reasons: (1) Only the phase being struck by lightning must be analyzed, because the voltages induced in the other phases will be much lower. (2) Assumptions about the lightning stroke are by necessity very crude, and very refined line models are therefore not warranted. (3) The risk of insulation failure in substations is highest for backflashovers at a distance of approx. 2 km or less. Insulation co-ordination studies are therefore usually made for nearby strokes. In that case, the modal waves of an M-phase line “stay 4-83 together”, because differences in wave velocity and distortion among the M waves are still small over such short distances. They can then easily be combined into one resultant wave on the struck phase. There seems to be some uncertainty, however, about the value of the surge impedance which should be used in such simplified single- phase representations. It appears that the “self surge impedance” Zs t-ourge Of Ba: (4-878) should be used. For nearby strokes it ts also permissible to ignore the series resistance. Attenuation ed by corona may be more important than that caused by conductor losses. At the time of writing these notes, corona is still diffi- cult to model, and it may therefore be best to ignore losses altogether to be on the safe side. 4.2.2.3 Mephase Lossless Line with Constant L and C' Additional explanations are needed for extending the method of Section 1 from single-phase lines to M-phase lines. In principle, the equations are first written down in the mdal domain, where the coupled M-phase line appears as if it consisted of M single-phase lines. Since the solution for single-phase lines is already known, this is straightforward. For solving the line equations together with the rest of the network, which is always defined in pha to phase quantiti quantities, these modal equations must then be transformed as schematically indicated in Fig. 4.35. For simplicity, let us assume that the line has 3 phases. Then, with the notations from Fig. 4.35, each mde is described by an equation of the form of Eq. (1.6), or 1 aaCt) 7 3 Mya(®) + hse, 24 ( 4) 1 Lypanle) =F yl + Mappa? (4,106) 1 tyeezel®) "Ze Yel + Math ezelt to) 4-84 Linear » att) Transform-|—°2B ICo- fr) , fr) p-°2C }-PHase-s| ~—— mooat pomaiIn-——>|—pHase-| DOMAIN DOMAIN Fig. 4.35 - Transformation between phase and modal domain on a three-phase line where each history term hist was computed and stored eerlier. For mode a, this history term would be hist, (t,) = - z Vogt) ~ tpg g(t) (4.107) and analogous for modes b and c. These history terms are calculated for both ends of the line as soon as the solution has been obtained at instant ¢, and entered into a table for use at a later time step. As indicated in Fig. 4,36, the history terms of a three-phase line would occupy 3 sections of the history tables for modes a, b, c, and the length of each section would be “uncreased! “8H “increased DOANE the travel, tine of the particular aode increased to the nearest integer multiple of At “. Since the modal travel times t, t> % differ from each other, the 3 sections in this table are generally of different length. This is also the reason for storing history terms as nodal values, because one has to go back different travel times for each mode in picking up history teras. For the solution at tine t, the *) A single-phase line would simply occupy one section, whereas a six-phase line would occupy six sections in this table. 4-85 history terms of Eq. (4.106) are obtained by using linear interpolation on the top two entries of each mode section. After the solution in each time etep, the entries in the tables of Fig. 4.36 must be shifted upwards by one location, thereby throwing away the Table Table hist, starting adérese | for mode a of a line oldest 2 values lt~s, used for inter~ patho! polation : t-2at enter oa —e newest values Fig. 4.36 - Table for history terms of values at the oldest point at tty. 0s sseq* for km for hist, mk “| a 1 a mode c transmission lines This is then followed by entering the newly calculated history terms hist(t) at the newest point t. 4-86 Ins| d of physically shifting values, the EMIP moves the pointer for the starting address of each section down by 1 location, When this pointer chee the end of the table, it then goes back again to the beginning of the table (“wrap-around table”) [91]). The initial values for the history terms must be known for t=0, At, -2at, The necessity for knowing them beyond t=0 comes from ‘increased” the fact that only terminal conditions are recorded. If the conditions were also given along the line at travel time increments of At, then the initial values at t=0 would suffice. For zero initial conditions, the history table is simply preset to zero. For linear ac steady-state conditions (at one frequency u), the history terms are first computed as phasors (peak, not rms), (4.108) 1 BIST, =~ Vy ~ Taye where V, and I,, are the voltage and current phasors at line end m (analogous for HIST,,). With HIST = |aIST| + eJ% the instantaneous history terms are then hist, (et) = [HIST, + cos(ut + o), with t#0, -at, -2ar, ... (4,108b) ‘a! Eq. (4.108) 18 used for single-phase lines as well as for M-phase lines, except that mode rather than phase quantities must be used in the latter case. Eq. (4.106) are interfaced with the rest of the network by transforming them from modal to phase quantities with Eq. (4.78a), [epee] = Prsccue | Deer] + [oeeate] . G10 with the surge admittance matrix in phase quantities, 4-87 zo = oO ate (4.109) [eure] * 2 b (41098) 0 0 et and the history tems in phase quantities, hist ‘la-2a phase]. [ossettg “| [c] | ete | + (4.1096) hist, ‘Le-2e For a lossless Line with constant L'and C', the transformation matrix [7,] will alvays be real, as explained in the last paragraph of Section 4.1.5.2. Tt is found as the eigenvector matrix of the product [c'][L"] for each particular tower configuration, where [L'] and (c'] are the per unit length series inductance and shunt capacitance matrices of the line. For balanced lines, (ty) is known a priori from Eq. (4.58), and for identical balanced three-phase lines with zero sequence coupling only it is known a priori from Eqs (4.65) The inclusion of Eq. (4.109) into the system of nodal equations (1.84) for the entire network is quite straightforward. Assuue that for the example of Fig. 4.35, rows and columns for nodes 1A, 1B, 1¢ follow each other, as do those for nodes 2A, 2B, 2¢ (Fig. 4.37). Then the 3x3 matrix [Y.44] enters into two 3 x 3 blocks on the diagonal, as indicated in Fig. 4.37, while the history tems [histPM@**] = (histy, 94+ Eq. (4.109¢) enter into rows IA, 1B, 1C, on the right-hand side with negative Esty pop? MStye 26] of signs. Analogous history terms for terminal 2 enter into rows 2A, 2B, 2C on the right-hand side. While [¥,..,] 18 entered into [6] only once outside the time-step loop, the history terms must be added to the right-hand side in each tine step. M-phase lossless line models are useful, anong other things, for (a) simple studies where one wants to investigate basic phenomena, 4-88 (>) in lightning surge studies, where single-phase models are no longer adequate, and ~ histyy on ~ hist, op > histy 96 ~ histo, ay ~ histy, 15 + histone a Io) Wi [il-thist) Pig. 4.37 - Entries for a three-phase line into system of equations (c) a8 a basis for more sophisticated models discussed later. Lightning surge studies cannot always be done with single-phase models. For simulating backflashovers on lines with ground wires, for example, the ground wire and at least the struck phase must be modelled (“2-phase line”). Since it 1s not always known which phase will be struck by the backflashover, it 4s probably best to model all three phases in such a situation ("4~phase line”). An example for such a study is discussed in Section 4.1.5.2, with 4~ phase lossless line models representing the distribution line, and single~ Not included in the data phase lo listing are switches (or some other elements) for the simulation of potential less line models representing the tower: flashovers from the tower top (nodes D) to phases A,B,C. 4-89 4.2.2.4 Single and M-Phase Distortionless Lines with Constant Parameters Distortionless line models are seldom used, because wave propagation on power transmission lines is far from distortionless. They have been implemented in the EMTP, nonetheless, simply because it takes only a minor modification to change the lossless line equation into the distortionless Line equation. A single-phase transmission line, or a mode of an M-phase line, is distortionless if R-f. (4.110) Losses are incurred in the series resistance R' as well as in the shunt conductance G'. The real shunt conductance of an overhead line is very small (close to zero), however. If its value must be artificially increased to make the Line distortionless, with a resulting increase in shunt losses, then it 4s best to compensate for that by reducing the series resistance losses. The EMTP does this automatically by regarding the input value Riyoye as an ‘INPUT indicator for the total losses, and uses only half of it for R', Rot ee ice ‘i (4.1) With this formula, the ac steady-state results are practically identical for the line being modelled as distortionless or with R lumped in 3 places; the transient response differs mainly in the initial rate of rise. From Bq. (4,111), the attenuation constant a becomes R inpur /[@ i a (4,112) The factor 1/2 can also be justified by using an approximate expression for the attenuation constant for lines with low attenuation and low distortion (48, p. 257], R s inpur /G* INPUT a eee (4.113) 4-90 which is reasonably accurate if R' << aL" and G' << ul'. This condition is fulfilled on overhead lines, except at very low frequencies. Eq. (4.112) is then obtained by dropping the term with G'iypyp and by ignoring the fact that the waves are not only attenuated but distorted as well. If @ user wants to represent a truly distortionless line where G' is indeed nonzero, then the factor 1/2 should of course not be used. The factor 1/2 4s built into the EMTP, however, and the user must therefore specify R' iypyr twice as large as the true series resistance in this cas With a known, an attenuation factor eis calculated (£= length of Line). The lossless line of Eq. (1.6) 48 then changed into a distortionless Line by simply multiplying the history term of Eq. (1.66) with this attenuation factor, hist (tO = EE v(ed = a(n] +e (4.114) The surge impedance remains the same, nanely Ji*/C. For M-phase lines, any of the M modes can be specified as Mixing 1s allowed (e.g., mode 1 could be modelled with lumped resistances, and modes 2 and 3 as distortionless). distortionle: Better results are usually obtained with the lumped resistance model described next, even though lumping of resistances in a few places is obviously an approximation, whereas the distortionless line is solved exactly if the travel time is an integer multiple of ar. 4.2.2.5 Single and M-Phase Lines with Lumped Resistances Experience has shown that a lossy line with series resistance R’ and negligible shunt conductance can be modelled with reasonable accuracy as one or more sections of lossless lines with lumped resistances in between. The simplest such approach is one lossless line with two lumped resistances R/2 at both ends. The equation for this model is easily derived from the cascade connection of R/2 - lossless line - R/2, and leads to a form which is identical with that of Eq. (1.6), 491 c= 1 po ye + teeter, (4.115) Twodi fied bad except that the values for the surge impedance and history terms are slightly modified. With Z, R and t calculated from Eq. (4.99), R woaitiea 2+ 7 and 1 ‘modified hist (t-1) = (ed + 2- Piero). This model with R/2 at both ends is not used in the EMIP. Instead, the EMIP goes one step further and lumps resistances in 3 places, namely R/4 at both ends and R/2 in the middle, as shown in Fig. 4.27. This approach was taken because the form of the equation still remains the same as in Eq. (4.115), except that R Zuodified 7 +5 4.116) now. The history term becomes more complicated”), and contains conditions from both ends of the line at tt, hist, (t= = Pate + 2 - aye v] 2 2ipodified -— BMA ceo t-Bagee] Gun 2 2 nodified Users who want to lump resistances in more than 3 places can do so with the built-in three-resistance model, by simply subdividing the line into shorter segmente in the input data. For example, 32 segments would produce lumped resistances in 65 places. Interestingly enough, the results do not change much if the number of lumped resistances is increased as long as R< the equation at the bottom of p. 391, left column, in [50] contains an error. I, and Ip should not be computed from Eq. (7b); instead, use Ty = -(/2)e,(e-t) ~biy (Et) with the notation of [50], where Z is modified Of Ea- (4.116). For Ip, exchange subscripts k and m. 4-92 For example, results in Fig. 4.30 for the distributed-parameter case were practically identical for lumped resistances in 3, 65 or 301 places. Fig. 4.29 shows as well that TNA results are closely matched with R lumped in 3 places only. One word of caution is in order, however. The lumped resistance model gives reasonable answers only if R/4 << 2, and should therefore not be used if the resistance is high. High resistances do appear in lightning surge studies {f the parameters are calculated at a high frequency, e.g., at 400 kiz in Table 4.25, where R' = 597.4 Q/km in the zero sequence mode. Lumping R in 3 places would still be reasonable in the case discussed there where each tower span of 90 m is modelled as one line, since 13.4 is still reasonably small compared with 2 = 1028 % If it were used to model a longer line, say 90 km, then R/4 = 134009, which would produce totally erroneous results”), In such a situation it might be best to ignore R altogether, or to use the frequency-dependent option if higher accuracy is required. For Mephase lines, any of the M modes can be specified with lumped Tesistances. Mixing is allowed (e.g., mode 1 could be modelled with lumped resistances, and modes 2, ... Mas distortionless). The lumped resistances do not appear explicitly as branches, but are built into Eq. (4.115), (44116) and (4,117) for each mode, Should a user want to add them explicitly as branches, e.g., for testing purposes, then they would have to be specified as M xM~ matrices [R] in phase quantities, which could easily be done with the M-phase nominal recircuit input option by setting L=0 and C=O. All modes would have to use the lumped resistance model in this set-up, that is, mixing of models would not be allowed in it. "The UBC version of the EMIP stops with an error message if R/4 > Z. It would be advisable to add a warning message as well as soon as R/4 gets fairly large (e.g., > 0.05 * 2). 4-93 4.2.2.6 Single and M-Phase Lines with Frequency-Dependent Parameters The two important parameters for wave propagation are the characteristic impedance . Se too Voaae (4.118) and the propagation constant ys HLTH (4119) Both parameters are functions of frequency, even for constant distributed parameters R', L', G', C' (except for lossless and distortionless lines). The line model with frequency-dependent parameters can handle this case of constant distributed parameters”), even though it has primarily been developed for frequency-dependent series impedance parameters R'(u) and L'(w). This frequency-dependence of the resistance and inductance is mst pronounced in the zero sequence mode, as seen in Fig. 4.20. Frequency dependent line models are therefore important for types of transients which contain appreciable zero sequence voltages and currents. One such type is the single lne-to-ground fault. To develop a line mdel with frequency-dependent parameters which fits nicely into the EMTP, it is best to use an approach which retains the basic idea behind Bergeron's method. Let us therefore look at what the expression v + 24 used by Bergeron looks like now, as one travels down the line. Since the parameters are given as functions of frequency, this expression must first be derived in the frequency domain. At any frequency, the exact ac steady-state solution is described by the equivalent rcireuit of Eq (1.13), or in an input-output relationship form more convenient here, Vy cosh(ye) 2, sinh v8)"] Tv, -{t (4.120) la Fr stam yg) cosh ys) 1, c “mis case differs from the line with lumped resistances inasmuch as the resistance becomes truly distributed now. 4-94, which can be found in any textbook on transmission lines. Assune that we want to travel with the wave from node m to node k. Then the expression V + 2,1 is obtained by subtracting Z, times the second row from the first row in Eq. (4.120), = sett e We - Alyy 2 Vy t Zl er, (4-121a) or rewritten as Tm 7 Vgl2q ~ Wyle + Ty) 2 ot (4-121) with a negative sign on I,, since its direction is opposite to the travel direction. Eq. (4-121) 18 very similar to Bergeron's method; the expression V + 2,1 encountered when leaving node m, after having been multiplied with « propagation factor e'*, 4s the same when arriving at node k. This is very similar to Bergeron's equation for the distortionless line, except that the factor ts e“ there, and that Eq. (4.121) is im the frequency domain here rather than in the time domain. Before proceeding further, it may be worthwhile to look at the relationship between the equations in the frequency and time domain for the simple case of a lossless line. In that case, 2,0 [ery wiOer, and oe ot Anybody familiar with Fourier transformation methods for transforming an equation from the frequency into the time domain will recall that a phase shift of e 5" in the frequency domain will become a time delay + in the time domain. Furthermore, Z, is now just a constant (independent of frequency), and Eq. (4.121) therefore transforms to vt) ~ ZA a(t) = vat) + Zt (tor) which is indeed Bergeron's equation (1.6). For the general lossy case, the propagation factor Aw) = eM = oH. SPE 4-95 with y = a+ 38, contains an attenuation factor e™ as well as a phase shift e744, which are both functions of frequency. To explain its physical meaning, let us connect a voltage source V, to the sending end m through source @ source impedance which {s equal to Z.(w), to avoid reflections in m (Fig. 4.38). In - + Furthermore, let us 38). In that case, Vi + ZT = Vecuece? Furth » let us assume that 2(w) x n Tk “source Fig. 4.38 - Voltage source connected to end m through matching impedance the receiving end k is open. Then from Eq. (4-121), v ke” Yeouree * AC? (4.122) that is, the propagation factor is the ratio (receiving end voltage)/(source voltage) of an open-ended line if the line 1s fed through a matching * impedance Z_(w) to avoid reflections at the sending end. If Vj... = 1-0 at all frequencies from de to infinity, then its time domain transform Veource(t) Would be a unit impulse (infinitely high spike which is infinitely row with an area of 1-0), and the integral of Veoureg(t) would be a unit step. Setting Veouroe * 1-0 in Eq. (4.122) shows that A(u) transformed to the time domain mist Se the impulse which arrives at the other end k, if the *) one could also connect a matching impedance Z,(w) from node k to ground to avoid reflections at the receiving end as well. In that case, the left hang side of Eq. (4.122) becomes 2V, rather than Vy. Note that the ratio eV starts from 1.0 and becomes less than 1.0 as the line length (or frequency) is increased. This is in contrast to the open-circuit response V/Vq * 1-O/cosh(y£) more familiar to power engineers, which increases with"length or frequency (Ferranti rise). 4-96 source is a unit impulse. This response to the unit impulse, a(t) = inverse Fourier transform of {A(w)} (4.123) will be attenuated (no longer infinitely high), and distorted (no longer infinitely narrow). Fig. 4.39 shows these responses for a typical 500 kV aw aw ace 6000 420000 400000) t 1 300000 300000} 4000 200000 200000} 2000. 100000 00000] ° ° Fx 7.0 ° 10 0.50 0.58 © (me) ©) + as) nin inex (a) zero sequence mode (b) positive sequence (c) positive sequence mode with same mode with scale as (a) expanded scale Fig. 4.39 - Receiving end response v,(t) = a(t) for the network of Fig. 4.38 £ Vecurce(t) = uit impulse [94] . Reprinted by permission of Je Marti line of 100 miles length. They were obtained from the inverse Fourier trans- formation of A(w) = exp(-y£) calculated by the LINE CONSTANTS supporting routine at a sufficient number of points in the frequency domain. The ampli- tude of the propagation factors A(w) for the case of Fig. 4.39 is shown in Fig. 4.40. The unit impulse response of a lossless line would be a unit impulse at tet with an area of 1.0. In Bergeron's method, this implies picking up the history term v./2 + i, at tt with a weight of 1.0. In the more general 4-97 case here, history terms must now be picked up at more than one point, and aw) aw) 1.0 10 os os. ° ° 107? 2 ao? —+ 208 1? 2 10? —> 108 # (He) £ iz) (a) zero sequence mode (b) positive sequence mode Fig. 4.40 - Propagation factor A(u) for the line of Fig. 4.39 [94] - Reprinted by permission of J. Marti weighted with the “weighting function” a(t). For the example of Fig. 4.39(a), history terms must be picked up starting at 1,4, = 0-6 ms back in time, to approx. t,,, "2-0 ms back in time, The value t,,, i the travel time of the fastest waves, while t,,, is the travel time of the slowest waves. Each term has its own weight, with the highest weight of approx. 5400 around t= 0.7 ms back in time, Mathematically, this weighting of history at the other end of the line is done with the convolution integral hist, opagation™ ~ { tereocar‘t¥acw)au (4.124) Tain which can either be evaluated point by point, or more efficiently with recursive convolution as discussed later, The expression 1, .¢,, in Ea- (4.124) 46 the sum of the line current 4.) and of a current which would flow through the characteristic impedance if the voltage v, were applied to it (expression I,, + V,/2, in the frequency domain). 4-98 paaey *p Jo voyssyuxod Aq poauradey “[y6] 6E°7 “BFA 3 SUET aya 205 (m)°Z souEpaduy OfIBFXeR00IEUD - Ty"y “FTA ‘apou eousnbes eatatsod (q) (an) 3 — got oot T ‘0z- Ot- 0 ot oz 0 (2H) 3 + got ot T ot o0z 00€ oov 005 009 002, (see262p) (m)°z zo at6ue aseua (w) (72 30 apnatubeH (2H) 5 pou eousnbas oxez (v) ot 0b 00s 009 oz 008 {seexSep) (m)°z 30 eTéue aseua (y) (mz 30 epnzTubeH 4-99 With propagation of the conditions from m to k being taken care of through Eq. (4.124), the only unresolved issue is the representation of the term V,/Z, in Eq. (4.121b). For the same 500 kV line used in Fig. 4.39, the magnitude and angle of the characteristic impedance 2, are shown in Fig. 4.41, If the shunt conductance per unit length G’ were ignored, as is usual- ly done, 7, would become infinite at uO. This complicates the mathematics somewhat, and since G’ is not completely zero anyhow, it was therefore deci~ ded to use a nonzero value, with a default option of 0.03 us/km. As origin ally suggested by E. Groschupf [96] and further developed by J. Marti [94], such a frequency-dependent impedance can be approximated with a Foster-I R-C network, Then the line seen from node k becomes a simple R-C network in propagation (F18+ 4+42(a))+ One can then apply the trapezoidal rule of integration to the capacitances, or use any parallel with a current source hist other method of implictt integration. This transforms each R-C block into @ current source in parallel with an equivalent resistance, or into a voltage source in series with an equivalent resistance. Summing these for all R-C Dlocks produces one voltage source in series with one equivalent resistance, or one current source in parallel with one equivalent resistance (Fig. 4.42- ())+ In the solution of the entire network with Eq. (1.8), the frequency- dependent line is then simply represented again as a constant resistance Requiy t° Stound, in parallel with @ current source beta, + MSt as saetont which has exactly the same form as the equivalent circuit for the lossless line. To represent the line in the form of Fig. 4.42 in the EMTP, it is neces sary to convert the line parameters into a weighting function a(t) and into an R-C network which approximates the characteristic impedance. To do this, 2, and y are first calculated with the support routine LINE CONSTANTS, froa de to such a high frequency where both A(w) = exp(-y2) becomes negligibly small and 2 (w) becomes practically constant. J. Marti [94] has shown that At 4s best to approximate Aw) and Z,(w) by rational functions directly in the frequency domain. The weighting function a(t) can then be written down explicitly as a sun of exponentials, without any need for numerical inverse Fourier transformation. Similarly, the rational function approximation of 2,(u) produces directly the values of R and C in the R-C network in Fig. 4h 4-100 Ristac Mstpyopagation i set CU i= hist, ji ‘Propagation (a) with R-C network (>) with equivalent resistance after applying implicit integration Fig. 4.42 - Frequency-dependent line representation seen from line end k The rational function which approximates A(u) has the form ee cg cee eens (4.25) ‘approx CaFp)(aFpg) CoA) with 6 = jw and nm. The subscript “approx” indicates that Eq. (4.125) is strictly speaking only an approximation to the given function A(w), even = jut ‘mis though the approximation is very good. The factor e ™ {8 included to take care of the fact that a(t) in Fig. 4.39 4s zero up to ter, 5 this avoids fitting exponentials through the portion 0 < t < 1,4, where the values enber that a tine shift -t in the time domain is a phase ate zero anyhow (1 shite J*T in the frequency donain). Al poles py and zeros 2, in Eq. (4.125) are negative, real and simple (multiplicity one). With ntaane =0 fort < Tain" (4,127) This weighting function a, .4,' hist of Eq. (4.124) in each time step. Because of its form as a propagation sun of exponentials, the integral can be found with recursive convolution much more efficiently than with a point-by-point integration. If we look at Py (Ten) (t) 18 used to calculate the history term the contribution of one exponential term k, © mp, (t-t 1) s(t) =f iCewkye 7 MP ao (4.128) 5 ‘ain then 8,(t) can be directly obtained from the value s,(t-At) known from the preceding time step, with only 3 multiplications and 3 additions, CO) cys ay(tr ae) te, + Merny.) +e, + Mert cat), (4,129) 1 as explained in Appendix V, with c,,c),c, being constants which depend on the particular type of interpolation used for 1. The characteristic impedance Z,(u) is approximated by a rational function of the form [94] (tz, )(et2,)+00(stz.) Zemappron ®) ~ * Carp, Warp, )oosCSFE, (4.130) 4-102 with s=jo . All poles and zeros are again real, negative and simple, but the number of poles is equal to the number of zeros now. This can be expressed 2a approxt®) * : (4.31) whtch corresponds to the RC network of Fig. 4.42, vith Rok ° : (4.132) k age by ete. Rather than applying the trapezoidal rule to the capacitances in Fig. 4-42, J. Marti chose to use implicit integration with Eq. (1.3) of Appendix I’), with linear interpolation on i. For each R-C block . av a 1 trata ae Which has the exact solution ~a,at fay (tu) + vy (erat) + e (udu, (4,133) vy(e) =e z i othe with ay#1/(R,C,)- By using Linear interpolation on 1, the solution takes the form of v(t) = R, equivet * H(t) + ey (tHe), (4.134) with e,(t-At) being known values of the preceding time step (formula omitted for simplicity), or after suaming up over all R-C blocks and R, VCE) ™ Reguiy + H(t) + e(erbe) (4.1354) with *)tis method is identical to the recursive convolution of Appendix V applied to Eq. (4.131). Whether recursive convolution is better than the trapezoidal rule is still unclear. 4-103 a a R, =“R +) R, and e=)e,, (4.135b) equiv "8 * L Requiv-t Hes which can be rewritten as a(t) = V(t) + histye + (4.136) ‘equiv The equivalent resistance Roos, the sum of the history terms histy + Met. eatiog enters into the right hand side, The key to the success of this approach is the quality of the rational enters into matrix [G] of Eq. (1.8), whereas function approximations for A(a) and Z(w). J. Marti uses Bode's procedure for approximating the magnitudes of the functions. Since the rational functions have no zeros in the right-hand side of the complex plane, the corresponding phase functions are uniquely determined from the magnitude functions (the rational functions are minimum phase-shift approximations in this case) [94]. To illustrate Bode's procedure, assume that the magnitude of the characteristic impedance in decibels is plotted as a function of the logarithm of the frequency, as shown in Fig. 4.43 [94]. The basic principle 4s to approximate the given curve by straight-line segments which are either horizontal or have a slope which is a multiple of 20 decibels/decade. The points where the slopes change define the poles and zeros of the rational function. By taking the logarithm on both sides of Eq. (4.130), and multiplying by 20 to follow the convention of working with decibels, we obtain 20 log |z, (s)| = 20 log k + 20 log|stz,|...+ 20 log |stz,| ~ 20 log |stp,{...-2010g |s+p,1 (4.137) ‘c~approx' For s = Ju each one of the terns in this expression has a straight-Line asyuptotic behaviour with respect to w For instance, 20 log |Jw+ 2,| becones 20 log 2, for w << f,, which 18 constant, and 20 log w for w>> 2), which is a straight line with a slope of 20 db/decade. The approximation to Eq. (4.137) is constructed step by step: Each time a zero corner (at wrz,) is added, the slope of the asymptotic curve 1s increased by 20 db, or decreased by 20 db each time a pole corner (at w= p,) is added. The rt 4-104 zw) (2B) a0. a) 10° 10 £ (Hz) Fig. 4.43 - Asymptotic approxination of the magnitude of Z,(w) [94]. Reprinted by permission of J. Marti straight-line segments in Fig. 4.43 are only asymptotic traces; the actual function becomes a saooth curve without sharp corners. Since the entire curve is traced from de to the highest frequency at which the approximated function becomes practically constant, the entire frequency range is approxi- mated quite closely, with the number of poles and zeros not determined a priori. J. Marti improves the accuracy further by shifting the location of the poles and zeros about their first positions. Fig. 4.44 shows the magni- tude and phase errors of the approximation of A(w), and Fig. 4.45 shows the errors for the approximation of Z,(u) for the line used in Fig. 4.39. L. Marti has recently shown [95] that very good results can be obtained by using lower-order approximations with typically 5 poles and zeros rather than the 15 poles and zeros used in Fig. 4.44 and 4.45. Furthermore, he shows that positive and zero sequence parameters at power frequency (50 or 60 Hz) can be used to infer what the tower geometry of the line was, and use this geometry in turn to generate frequency-dependent parameters. With this approach, simple input data (60 Hz parameters) can be used to generate a 4-105 Magnitude error of ACs) Magnitude error of A(w) 207 a 10? — a0® ; £2) £ =) i i i goa i 3 eat = 22 Ry t i Boo i Ea é z- a 1 10? — 0"? a 10? —+ 10% £ (80) a) (a) zero sequence node (b) positive sequence mode (15 zeros and 20 poles) (13 zeros and 17 poles) Fig. 4.44 - Errors in approximation of A(w) for line of Fig. 4.39 [94]. Reprinted by permission of J. Marti frequency-dependent line model internally which is fairly accurate. For M-phase lines, any of the M modes can be specified as frequency- dependent, or with lumped resistances, or as distortionless. Mixing is allowed. A word of caution is in order here, however: At the time of writing these notes, the frequency-dependent line model works only reliably for balanced line: For untransposed line approximate real and constant 4-106 o z Goa a a . % 2 | po 4 d- B Bo i i bs Te a » bad 2 wea ao? — 108 £ ; 3 £2) & § i i go, $4 eal 31 3!, : S 6 So é é go. f : : i -« E+ 2 10? —+ 108 a yo? — uf # oe) fH) (a) zero sequence mode (b) positive sequence mode (15 zeros and poles) (16 zeros and poles) Fig. 4.45 - Errors in approximation of Z,() for line of Fig. 4.39 [94]. Reprinted by permission of J. Marti transformation matrices must be used, as explained in Section 4.1.5.3, which seems to produce reasonably accurate results for single-circuit lines, but 4-107 not for double-circuit lines. Research by L. Marti into frequency-dependent transformation matrices in connection with models for underground cables will hopefully improve this unsatisfactory state of affaire. Field test results for a single-line-to-ground fault from Bonneville Power Administration have been used by various authors to demonstrate the accuracy of frequency-dependent line models [84]. Fig. 4.46 conpares the field test results with simulation results from an older method which used two weighting functions a, and a, [84], and from the newer method described here. The peak overvoltage in the field test was 1.60 p.u., compared with 1.77 p.u. in the older method and 1.71 p.u. in the newer method. Constant 60 Hz parameters would have produced an answer of 2.11 p.u. 4-108 312.5 kV ° 0.02 0.04 4 0.06 t (s) (») 312.5 kV ° 002 0.04, 0.06 t (s) ) Fig. 4.46 - Comparison between voltages at phase b for [94]: (a) Field test oscillograph (b) BPA's frequency-dependence simulation (c) New model simulation Reprinted by permission of J. Marti SL 5. UNDERGROUND CABLES There is such a large variety of cable designs on the market, that it is difficult, if not impossible, to develop one computer program which can calculate the parameters R', L', C' for any type of cable. For lower insulated with voltage ratings, the cables are usually unscreened and polyvinyl chloride. An example of a three-phase 1 kV cable with neutral conductor and armour is shown in Fig. 5.1. Fig. 5.1 - Armoured 1 kV cable (I=stranded conductor, 2 = insulation, Fig. 5.2 - 3ebedding, 4 = flat steel wire armour, 5 = helical steel tape, 6 = plastic outer sheath). Reprinted by permission from Siemens Catalog 1980 12 to 35 kV distribution cable with concentric neutral conductors (1 = stranded conductor, 2, 4 = conductive layers, 3 = plastic insulation, 5 = conductive tape, 6 = concentric neutral conductors, 7 = helical copper tape, 10 = inner sheath, 11 = plastic outer sheath). Reprinted by permission from Siemens Catalog 1980 At the distribution voltage level, the cables are usually screened with concentric neutral conductors, as shown in Fig. 5.2. At the transmission voltage level, two types of cables are in widespread use today, namely the pipe-type cable (Fig. 5.3) and the self-contained cable (Fig. 5.4). In the pipe-type cable, three paper-insulated ofl-impregnated Steet pine (tite th rating oi ‘Skid wires Metol topes Paper it insulation Screen: Conductor (stranded coppery Fig. 5.3 - Pipe-type oil-filled cable [148]. © 1979 John Wiley & Sons, Ltd. Reprinted by permission of John Wiley & Sons, Ltd cables are drawn into a steel pipe at the construction site. The helical skid wires make it easier to pull the cables. After evacuation, the pipe is filled with ofl and pressurized to a high pressure of approx. 1.5 kPa. Pipe-type cables are used for voltages from 69 to 345 kV, with 550 kV cables under development. The typical self-contained oil-filled cable is a single-core cable (Fig- 5-4). Its stranded core conductor has a hollow duct which is filled with ofl and kept pressurized with low-pressure bellow-type expansion tanks. Underground and submarine self-contained cables are essentially identical, except that underground cables do not always have an armour. Gas-insulated systems with compressed SF, gas are used for compact substation designs. The busses in such substations consist of tubular conductors inside a metallic sheath, with the conductors held in place by plastic spacers at certain intervals (Fig. 5.5). SF,-busses are in use in lengths of up to 300 m. A similar design can be used for cables, but SF, - cables are still experimental, with the sheath usually being corrugated. In Fig. 5.4 - Single-core self-contained cable (C = stranded core conductor with oil-filled duct, I = paper insulation, S = metallic sheath, B = bedding, A = armour, P = plastic sheath). Details of conductive layers left out (a) Single-phase (b) Three-phase Fig. 5.5 - SF, bus EMTP studies, such relatively short busses can often be ignored, or representd as a lumped capacitance. Only in studies of fast transients with high frequencies mst SF,-busses be represented as transmission lines. Since the single-phase geometry is essentially similar to that of a self-contained cable, and since the three-phase geometry is similar to that of a pipe-type cable, no special programs are needed to handle SFg-busses or cables, except that the three-phase arrangement of Fig. 5.5(b) has no electrostatic screens as in the case of a pipe-type cable of Fig. 5.3. Fig. 5.1 to 5.5 are only a few examples for the large variety of cable designs. The support routine CABLE CONSTANTS was developed by A. Ametani essentially for the coaxial single-core cable design of Fig. 5.4 and 5.5(a), and later expanded for the pipe-type cable of Fig. 5.3 and for the three-phase SF,-busses of Fig. 5.5(b)- At this time, there 1s no support routine for the types of lower voltage cables shown in Fig. 5.1 and 5.2, but calculation methods applicable to such non-coaxial arrangements are briefly discussed in Section 5.7. 5.1 Single~Core Cables The cable parameters of coaxial arrangements, as in Fig. 5.4, are derived in the form of equations for coaxial loops [150, 152]. In Fig. 5.4, loop 1 is formed by the core conductor C and the metallic sheath S as return, loop 2 by the metallic sheath $ and metallic arnour A as return, and finally loop 3 by the armour A and either earth or sea water as return. Selel S The series impedances of the three loops are described by three coupled equations 5 4 a e | [a - te te} |e ow 0 2 a} |G The self impedance Z',, of loop 1 consists of 3 parts, 2'11* 2 core-out * 2" core/sheath-insulation * “'sheath-in, G2) with ers internal impedance (per unit length) of tubular core conductor with return path outside the tube (through sheath here), impedance (per unit length) of insulation between core and sheath, and ® core/ sheath-insulation ory = internal impedance (per unit length) of tubular sheath with return path inside the tube (through core conductor here). Similarly, 242 * **sheath-out *?"sheath/armour-insulation * ”’armour-in, (°°9) and 2 (6.4) 33 ~ ?"armour-out * ?'armour/earth-insulation * ”'earth, with analogous definitions as for Eq. (5.2). The coupling impedances Zp = 2"p1 and Z',, = 2'y9 are negative because of opposing current directions (I, in negative direction in loop 1, I in negative direction in loop 2), 2" =--2 (5.5a) 127 Ban sheath-mutual 243224) 2-2! 237 2"32 armour-mutual (5.5b) with 2" os cath-mutual 7 mtval impedance (per unit length) of tubular sheath between the inside loop 1 and the outside loop 2, and 2" srmour-mutual " Mtval impedance (per unit length) of tubular armour between the inside loop 2 and the outside loop 3. Finally, Z2',, = Z's; = 0 because loop 1 and loop 3 have no common branch. The simplest terms to calculate are the impedances of the insulation, which are simply B 2 gk so 3S ink (5-6) with y, = permeability of insulation, a insulation F = outside radius of insulation, + in identical units (e.g. in ma) 4 = inside radius of insulation, If the insulation is missing, e-g-, between armour and earth, then ?"tnsulation 7 ° ‘The internal impedance and the mitual impedance of a tubular conductor with inside radius q and outside radius r (Fig. 4.5) are a function of frequency, and are found with modified Bessel functions [149]. . = om m 7 2" becta rage (ma)K, (mr) + K, (maT, (mr) } G.7a) ' pm 2" ube-out by (1, (ar) Ky (aq) + K, (ar) 1, (mq)} G7) ' - 2 e 7" eube-mtual ” ZaqzD come with D = I,(mr) Ky(mg) - I,(mq) K, (mr). (5.74) The parameter a= VIO G.Te) is the reciprocal of the complex depth of penetration p defined earlier in Bq. (4.5) except that u, im Eq.(4.5) is replaced by w= uu, here, with u,#l if the conductor material is magnetic. A subroutine SKIN alcula 1 impedance 2! “6. subroutine SKIN for calculating the imped 2 eube-out Of Bae (5-76) was developed at BPA for the support routine LINE CONSTANTS, and later modified at UBC to “TUBE” for the calculation of 2" upon M4 2" eybe mutual as well. All argunents of the modified Bessel functions Tg, I), Kg, K, are complex numbers with a phase angle of 45 because of Eq. (5.7e). In such a case, the following real functions of a real variable can be used instead: ber(x) + jbel(x) = In(x YD, ber'(x) + sbei'GD = ¥ILYGYD, (5.8) ker(x) + jkei(x) = Kg(x/J), ker'(x) + jkei'(x) = -¥IK\ Gv). These functions are evaluated numerically with the polynomial approximations of Hq. (911-1) to (9-11.14) of [149]. For arguments x < 8, the absolute error is < 107, whereas for arguments x > 8, the relative error is < 3+10 °. To avoid too large numbers in the numerator and denominator for large arguments of x, the expressions f(x) and g(x) in Eq. (9.11.9) and (9.11.10) of [149] are multiplied with exp (- 1*4 vz have absolute values greater than 8, then in addition to the above x). If both arguments mq and mr multiplication, the Ky- and K,- functions are further mltiplied by exp (2mq) to avoid indefinite terms 0/0 for very large arguments. When the support routine CABLE CONSTANTS was developed, subroutine TUBE did not yet exist, and A. Ametani chose slightly different polynominal approximations for the functions I, I,, Ky, K, in Bq. (5-7). He uses Eq. (9.8.1) to (9.8.8) of [149] instead, with the accuracy being more or less the same as in the polynomials used in subroutine TUBE. Simpler formulas with hyperbolic cotangent functions in place of Eq. (5.7) were developed by M. Wedepohl [150], which also give fairly accurate answers as long as the condition (r-q)/(r+q) < 1/8 is fulfilled. This was verified by the author for the data of a 500 KV submarine cable. Peis tay (ore This is the earth or sea return impedance of a single buried cable, which is ‘The only term which still remains to be defined is z' discussed in more detail in Section 5.3. 5-8 Submarine cables always have an armour, while underground cables may only have a sheath. The armour often consists of spiralled steel wires, which can be treated as a tube of equal cross section with y,# 1, without too much error [153]. Eq. (5.1) is not yet in a form suitable for EMTP models, in which the voltages and currents of the core, sheath and armour mst appear, in place of A wore accurate representation is discussed in [151]. loop voltages and currents. the terminal conditions The transformation is achieved by introducing Veore ~ Veheath * Toore v. and Ly = Iyneaen * Leore (5.9) 3 * Yarmour Tarnour * Tsheath * Teore where V.j., 7 voltage from core to ground, Veheath " Voltage from sheath to ground, Varmour 7 Voltage from armour to ground. By adding row 2 and 3 of Eq. (5-1) to the first row, and by adding row 3 to the second row, we obtain Lax too Beg 2h core cc Zcs Zeal [Teore ~ [oneacn/@*] = ("ec "a5 sa} |Toheath (5.108) iv Jax} zt a a ae armour ac 2s 2 aal [Tarmour| eit | cr oe cee ect Beg Be 7 Bap t B99 * 93 t 2153 > Bea ac Beg * Mas * 223 * 233» Ge Bag 7 Bat Mag + Ui33 > Baa = 33° Some authors use equivalent circuits without mutual couplings, in place of the matrix representation elements) and mtual impedances (off-diagonal elements). shows the equivalent circuit armour, which is essentially of Eq. (5-10) with self impedances (diagonal For example, [150] of Fig. 5.6 for a single-core cable without the same as the TNA four-conductor representation of overhead lines in Fig. 4.28. 11 12 A098. core TT ~ 2 OH. sheath 21 + 20 CO earth Fig. 5-6 - Three conductor I-circuit suitable for TNA's 5-12 Shunt _Admittances For the current changes along the cable of Fig. 5.4, the loop equations are not coupled, = aty/ax = (G', + Jul"), = at, /ax = (Gt, + Jub" Voy Gu) = aL,/ax = (G', + JuC", DV, G', and C', are the shunt conductance and shunt capacitance per unit length for each insulation layer. If there is no insulation (e.g-, armour in direct contact with the earth), then replace Eq. (5.11) by v0 G12) The shunt capacitance of tubular insulation with inside radius q and outside radius r is or = Ot (19 ing with €) = absolute permittivity or dielectric constant of free space (ey defined in Bq. (4-22)) and ¢, = relative permittivity or relative dielectric constant of the insulation material. ‘Typical values for ¢, are shown in Table 5.1 [54]. 5-10 Table 5.1 ~ Relative permittivity and loss factor of insulation material [54]. Reprinted by permission of Springer-Verlag and the authors Relative Permittivity |Loss Factor tang Insulation Material at 20°C at 50 Hz and 20°C butyl rubber 3.0 to 4.0 0.05 insulating of] 2.2 to 2.8 0.001 to 0.002 ofl-impregnated paper 3.3 to 4.2 0.003 to 0.008 polyvinyl chloride 3.0 to 4.0 0.02 to 0.10 polyethylene 2.3 0.0002 crosslinked polyethylene 2.4 0.0008 | The shunt conductance G' is ignored in the support routine CABLE CONSTANTS, which 1s probably reasonable in most cases. It cannot be ignored, however, 1f buried pipelines are to be modelled as cables, as explained in Section 5.6. If values for G' are available for cables, it is normally in the form of @ dielectric loss angle 6 or loss factor tané. Then Gt = uC! + tand (5.14) Typical values for tané are shown in Table 5.1. In the literature on electromagnetics, the shunt conductance is usually included by assuming that ¢, in Eq. (5.13) is a'complex mumber €, = e' - je", with Eq. (5.13) rewritten as Gt + soct = RY Cer — Ger, (5.15) ne For cross-linked polyethylene, both ¢' and e” are more or less constant up to 100 Miz [201], with the typical values of Table 5.1. For oil-impregnated Paper insulation, both c’ and e” vary with frequency. Measured values between 10 kHz and 100 MHz [154] showed variations in e' of approximately 20%, whereas c” varied much more. Fig. 5.7 shows the variations which can be expressed as a function of frequency with the empirical formula 0.94 e, 725+ —_O%, G16) (a+ju6 1079) coe e" — £ (iz) Fig. 5.7 - Measured values of €' and” for a cable with of1-impregnated paper insulation at 20°C [154]. Reprinted by permission of TEE and the authors The support routine CABLE CONSTANTS now assumes 0 and c! being constant, but ie could easily be changed to include empirical formulas based on neasurenents, such as Hq. (5.16). At this time, formulas based on theory are not available because the frequency-dependent behaviour of dielectrics 1s too complicated. Except for very short pulses (<5 us), the dielectric losses are of Little importance for the attenuation [154], and using a constant ¢' with e" = 0 should therefore give reasonable answers in most cases. Again, Eq. (5.11) is not yet in a form suitable for EMTP models. With the conditions of Eq. (5-9), they are transformed to ‘core/@* ae ° core e - |e V 5.17 Teneath/ 5 fsneath| "7? al, Tt ° Y'y + ¥'31 [Varmour| 5-12 where ¥', = 6 5.2 Parallel Single-Core Cables ‘There are not many cases where single-core cables can be represented with single-phase models. A notable exception is the submarine cable system, where the individual cables are laid so far apart (to reduce the risk of anchors damaging more than one phase) that coupling between the phases can be ignored. In general, the three single-core cables of a three-phase underground installation are laid close together so that coupling between the phases mst be taken into account. If we start out with loop analysis, then it is apparent that it is only the most outer loops (armour with earth return, or sheath with earth return in the absence of armour) through which the phases become coupled. The magnetic field outside the cable produced by loop 1 and 2 in Fig. 5.4 is obviously zero, because the field created by I, in the core is exactly earth surface Fig. 5.8 - Three single-core cables cancelled by the returning current I, in the sheath, etc. The first two equations in (5.1) are therefore still valid, whereas the third equation now h coupling terms among the three phases a,b,c, or 5-13 oo ° ° o Toop : ° Ze | 6-18) Ze 2126 0 syametric Bare 2226 2'23¢] ° 2326 2'33¢! with Z' 4, Z',os Z'y¢ being the mitual impedances between the three outer loops of Fig. 5.8. By using Eq (5-9) for the transformation from loop to conductor (core, sheath, armour) quantities, the matrix in Eq.(5.18) becomes sert-a) — ("matuat and! [pstuar ee! 2" conal eerep) — eueuan we! | (5.19) syametric 2" cette! ‘The 3 x 3 submatrices [2,416 4] matrix in Eq. (5-10a) for each cable by itself, whereas the 3 x 3 off-diagonal matrices have identical elements, e-g-, ] etc. on the diagonal are identical to the ‘a Zap ad (2yucuar ab) * |2'ap Zab 2a) ) Ban May ap) The only elements not yet defined are the mitual impedances Z' 4, 2' 40s 2'yo S14 of the outer earth return loops, which are discussed in more detail in Section 5.3. If one of the cables does not have an armour, its self submatrix is obviously a 2 x 2 matrix and its mutual submatrix is a 2 x 3 matrix. For cables without sheath and armour, the submatrices become 1 x 1 and 1 x3, respectively. ‘There is no coupling anong the three phases in the shunt admittances- Therefore, the shunt admittance matrix for the three-phase system is simply yt] 0 0 epererls hOl ee az.0] 0 (5-21) o 0 where [Y",] is the 3 x 3 matrix of Eq. (5.17) for phase a, ete- The screening effect of the sheath and armour depends very mch on the method of grounding. For example, if cable a is operated at 100 A between core and ground, with sheath and armour ungrounded and open-circuited, then the full 100 A will flow in the outer loop (Loop currents I, = 100, I, = 100, 1, = 100 in Fig. 5.4). This will produce maximun induced voltages tn the conductors of @ netghbouring cable b. How much nuisance this induction effect creates depends again on the method of grounding within cable b itself. If cable b is operated between core and ground (loads connected from core to ground), and if its sheath and armour are ungrounded and open-circuited, then the induced voltage will drive a circulating current through the core, ground and load impedances. If cable b is operated between core and sheath (loads connected from core to sheath), then there will be no circulating current in that loop because, according to Eq. (5-20), the induced voltages are identical in core and sheath. There would be a circulating current through the sheath and armour in parallel with earth return if the sheath (and armour) is grounded at both ends. If both the sheath and armour in the current-carrying cable a are grounded at both ends, then the voltage induced in the conductors of the neighbouring cable b would be small. For the practical example of a 500 kV ac submarine cable at 60 Hz, 14% of the core current would return through the sheath, 87.8% through the armour, and only 5.6% through the outermost loop 5-15 with ground or sea water return. ‘The induction effect in neighbouring cables would then be only 5.6% compared to the case with ungrounded sheath and armour. The algebraic sum is larger than 100% because there are phase shifts = e348") = 87.8¢ 13179" | among the three currents (Io. 0.ty a 1, = 5.6¢7 386° earth » 7 * 5.3 Earth-Return Impedances' In Eq. (5.4), the impedance of the loop formed by the outermost tubular conductor and the earth (or sea water) as return path is needed. This shall be called the “self earth-return impedance”. For the matrix of Eq. (5-18), the “mutual earth-return impedance” Z',, between the loop formed by the outermost tubular conductor and the earth return path of cable 1, and the analogous loop of cable k, is needed as well. The four main methods of installing cables are as follows [148]: (a) The cable is laid directly in the soil, in a trench which is filled with a backfill consisting of either the original soil or of other material with lower or more stable thermal resistivity. (b) The cable is laid in ducts or troughs, usually of earthenware or concrete. (c) The cable is drawn into circular ducts or pipes, which allows additional cables to be installed without excavation. (d) The cable is installed, in air, e.g. in tunnels built for other Purposes. In cases (a), (b) and (c) the cable is clearly buried underground, and formulas for buried conductors must therefore be used. In case (a), the radius R of the outermost insulation is simply the outside radius of the cable. In cases (b) and (c) it should be the inside radius of the duct if the duct has a similar resistivity as the soil, or the outside radius if it is a very bad conductor, or possibly some average radius if it is neither a good nor a bad conductor. What to do in case (d) is somewhat unclear. *) ‘The assistance of N. Srivallipuranandan and L. Marti in research for this section is gratefully acknowledged. 5-16 Reasonable answers might be obtained by representing the tunnel with an equivalent circular cross section of radius R. Another alternative is to assume that the tunnel floor is the surface of the earth, and then use the earth-return impedance formula for overhead conductors. This would ignore current flows in the earth above the tunnel floor. 5.3-1 Buried Conductors in Semi-Infinite Earth Exact formulas for the self and mtual earth-return impedances of buried conductors were first derived by Pollaczek [29]. In these formula: the earth is treated as semi-infinite, extending from the surface downwards and Fig. 5.9 - Geometric configuration of two cables sideways to infinity. If the horizontal distance between cable i and cable k is x, and if cable 4 and k are buried at depth h and y, respectively (Fig. 5-9), then the mitual earth-return impedance is [150] 2 pa’ 2 eseuar "Fe Ky(adKympy +f BLD a) ery jaxpda} (5-22) >Ja], he can then use Carson's tafinite series or asymptotic expansion discussed in Section 4.1.1.1. Fig. 5.10 and 5.11 show the errors in Anetani's results from support routine CABLE CONSTANTS, as well as the errors of Wedepohl's approximate formulas [150] for self impedance, 5-18 ‘ wpm py aR 4 2 earth "Ge fin GR + ost ah} (5.24) and for mutual tapedance 2 7 pa yaa 2 2 eueuar 7 Fe (nL + 0.5 2 ma} (5.25) e 3 Resistance Error ( % ) Induetenes Error (% ) Fig. 5.10 - Relative errors in self earth-return impedance formulas for buried conductors (R = 48.4 mm, p = 100 gm) [168] . Reprinted by permission of N. Srivallipuranandan with y= 0.577215665 (Euler's constant), and 5-19 4 = sum of the depths of burial of the two conductors. Wedepohl's approximations are anazingly accurate up to 100 kiiz (error < 1%), and then becone less accurate as the frequency increases (25% error at 1 MHz) where the condition | < 0.25 or al < 0.25 is no longer fulfilled. yance Error ( % ) mctan Induetonee Error (% ) Fig. Relative error in mutual earth-return impedance formulas for buried conductors (d = 0.3 m, h = 0.75 m, y = 0.75 m) [168]. Reprinted by permission of N. Srivallipuranandan Semlyen has recently developed a very simple formula based on complex depth p = 1/m [156], analogous to Eq. (4.3) for the case of overhead lines. 5-20 For the self earth-return impedance, the formula is ; = ju 1 Qearth ~ Ze MR + aR D> (5-26) while a similar formula for the mitual impedance has not yet been found. The error of Eq. (5.26) is plotted in Fig. 5.10. Considering the extreme simplicity of this formula as compared to Pollaczek's formula, it is amazing to see how reasonable the results from this approximate formula are- 5.3.2 In sone cases, it my be reasonable to assune that the earth is infinite in all directions around the cable. This assumption can be made when the depth of penetration in the earth fearen G08 A Pearth_ (aa) (» 6.27) becomes much smaller than the depth of burial. For submarine cables, where p is typically 0.2 Qm, this is probably more or less true over the entire frequency range of interest, whereas for underground cables it would only be true above a few Milz or so. Bianchi and luont [151] have used this Infinite earth assumption to find the sea return impedance of submarine cables. The self earth-return impedance for infinite earth is easily obtained —— to infinity. ‘Then with = R, : a K (aR) earth ~ 7K KyGRY * ony The mitual earth-return impedance was derived in [168] as K (ma) 6.29) 2 oo mutual 2m RyR,Ky(mR,) Ky (mR, ) 5.3.3. Overhead Conductors If the cable is installed in air, or laid on the surface of the ground, 5-21 then the earth-return impedances are the same as those discussed for overhead lines in Section 4. The support routine CABLE CONSTANTS uses Carson's formula in that cas For a cable laid on the surface of the ground, the height 1s equal to R. Ametani has tried a epecial formula of Sunde for conductors on the surface of the ground, but the answere were found to be very oscillatory around the seemingly correct answer. Sunde's formla was therefore not implemented. 5.3.4 Mutual 1 There is inductive coupling between the loop of an overhead conductor with earth return and the loop of a buried conductor with earth return. The mutual impedance between these two loops is needed, for example, for studying the coupling effects in pipelines from overhead lines, as discussed in Section 5.6. This case was treated by Pollaczek as well, with 2 7 io i exp {-hla|-y/ a? a7} ‘mutual ~ 27 -- lel +S +a As in the case of buried conductors, Ametani uses an approximation for this exp (Jax)da (5.30) integral by replacing y Va? +a with yla|- with ehte approxtnation, the formula becoaes identical with Carson's equations, with the height of the buried conductor having « negative value. In conection with a pipeline study [158], it was verified that Carson's formula and Pollaczek's formula sive Adencical results at 60 ls. At higher frequencies, the differences would probably be sintlar to those show ia Fig. 5.11. 5.4 Pipe-Type Cable Compared to the geometry of the single-core cable of Fig. 5.4, the geometry of the pipe-type cable of Fig. 5.3 is more complicated. It is therefore more complicated to obtain the impedances of a pipe-type cable, mainly for two reasons: 5-22 (@) The single-core cables inside the pipe are not concentric with respect to the pipe. (b) The steel pipe is magnetic, and subject to current-dependent saturation effects. The analysis is somewhat simplified by the fact that the depth of penetration into the pipe is less than the pipe thickness at power frequency and above. At 60 Hz, it is 1.5 mm from Eq. (5-27), with typical values of 9 = 0.2+1076am and p,= 400, whereas a typical pipe thickness for a 230 kV cable is 6.4 mm. For transient studies with frequencies above power frequency, the pipe thickness can therefore be assumed to be infinite, or equivalently, the earth-return can be ignored. Table 5.2 shows the Table 5.2 - Harth-return current in a 230kV pipe-type cable for single-phase fault (uy, = 400) current in earth (percent of core current) 94.50 31.00 0.85 00 current returning in the earth for a single~phase-to-ground fault in a 230 kV pipe-type cable, with the pipe being in contact with the earth. To arrive at these values, it was assumed that the core of the faulted phase was in the centre of the pipe, and that the two unfaulted phases can be ignored. With these assumptions, the impedance formulas of Section 5.1 can be used. If the two unfaulted phases were included, the earth-return current would probably be even less because some current would return through the shield tapes and skid wires of the unfaulted phases. The relative permeability yu, influences the values of Table 5-2; with », = 50, 6% of the current would return through the earth at 60 Hz, or 0.02% with pw, = 1600. 5-23 au Re. Be: If the depth of penetration is less than the pipe thickness, then no voltage will be induced on the outside of the pipe (z' = 0 from pipe-mitual Hq. (5.7e)), and consequently, the loop current pipe/earth return will be practically zero. Tn that case, the pipe is the only return path. ‘The configuration is then essentially the same as that of three single-core cables in Fig. 5.8, except that the pipe replaces the earth as the return path. Ié we assume that each phase consists of three conductors (e-g-, core, shield tapes represented as sheath, skid wires represented as armour), then the loop impedance matrix is the same as in Eq. (5.18). Coupling will only exist among the three outermost loops of each armour (skid wires) with return through the pipe. What is needed then is a formula for the self impedances Zz Zz Z',,, of the loops formed by each armour (skid wires) and the 33a, 7'33b, 7'33¢ pipe, and a formula for the mtual impedances 2',, 2"), | ye, Ziq between two such loops. The support routine CABLE CONSTANTS finds these impedances with formulas first derived by Tegopoulos and Kriezis [159], and later used by Brown and Rocanora [160]. In these formulas it is assumed that the current is concentrated in an infinitesimally small filament at the centre of each single-core cable. This model can be applied to conductors of finite radius 4€ proximity effects are negligible, either because of symmetrical positioning within the pipe, or because the conductor radius is small compared to the distance to other conductors or the pipe wall. In pipe-type cables, neither condition is met since the conductors are relatively large and lie on the bottom of the pipe. The pipe-type cable impedances from CABLE CONSTANTS are therefore not completely accurate, but no better analytical models are available at this time. Brown and Rocamora, who proposed the formulas originally, recommend methods based on the subdivision into partial conductors discussed in Section 5.7, for more accurate impedance calculation [161]. Hopefully, a support routine based on the subdivision method will become available some day. 5-26 The self impedance Z',,,, etc. of the loop between the armour (skid wires) and the pipe consists again of three terms, as in Eq. (5.4). The first term Z' armour-out 28 the sane as in Eq. (5.7b), with the assumption that proximity effects can be ignored. The second term for the insulation becomes more complicated than Eq. (5.6), because of the eccentric geometry, 3 ' Po iD ay 2 snoutation ” Izy MRO = (qd I (9.31) with q, Ry and 4, defined in Fig. 5-12. The third term for the internal Fig. 5.12 - Geometry of pipe-type cable (q,r = inside and outside radius of pipe; Ri,R, = outside radius of single-core cables i,k} 4; .4, = offset trom centre) impedance of the pipe, with return on the inside, replaces 2" (3.4): 5-25, . uote? ay 20 K,(a9) B yaperin ” 38 oe lemma LT) area = aye | O52) with m from Eq. (5-7e), and » = uu, = permeability of the pipe, K, = modified Bessel function of the second kind of order 4, Ki, = derivative of Ky. For the concentric case with d, = 0, Eq. (5.32) becomes identical with Eq. (5-28). The mitual impedance 2',,, etc. between two outermost loops formed by armour (skid wires) and pipe is Bo Beseuat ~ 1° Te - a Ko(mq) + piv ee by mak] Caay ay? + 4)? ~ 2d,d,cos0,,. 2 K(aq) 1 cos! 2 2 -= 5233) (0, (2n,. aa) = a a al (5-33) Except for replacing 2" cacey WEN 2" py ogitqr ANd FOF USING 2 ayeygy FFM Eq. (5-33) instead of (5-22), all calculations remain the same as in Section 5.2, including the transformation from loop to phase quantities. If the cables inside the pipe do not have an armour (skid wires), or a sheath (shield tapes), then some of the matrices will be reduced to 2x2, or 1x1, as discussed in Section 5.2. In practice, the shield tapes and skid wires can probably be represented as one single sheath. ‘The magnetic properties of the steel pipe are easily taken into account by using the proper values for the relative permeability , in Eq. (5.32) and (5.33). Unfortunately, u, depends on the current because of saturation effects, as shown in Fig- 5.13 [192]. To model saturation effects accurately is not simple, because even at one frequency, say at 60 Hz, the permeability would not remain constant over one cycle. A two-slope saturation curve was tried in [161], with the conclusion that reasonably accurate answers can be obtained with a constant value of u,+ ‘The sensitivity of the results with respect to u, can then be checked by re-running the case with one or nore different values of u,+ 5-26 —+ 1,/D, (A/mm) PoP Fig. 5.13 - Relative permeability as a function of pipe current ay current, D, = pipe diameter) [192]. © 1964 IEEE pipe Since the shield tapes and skid wires are in contact with the pipe wall, the values of the capacitances between the shield tapes/skid wires of the three phases and between them and the pipe are immaterial. They are shorted out. Eq. (5-21) can therefore be used directly for the shunt admittance matrix. The support routine CABLE CONSTANTS does not assume this contact with the pipe in the beginning, however, and is therefore more general. For this general case, a potential coefficient matrix is found first, Prag) [Pap] Pac a ey) +] a) By) Bye (5-34) cond TH} |B teq) toy) Bee) where [P',], (P',], (P',] are the 3x3 matrices of each single-core cable found by inversion of Eq. (5-17) with G’ = 0, ¢," -c," ° =r cect ct (5.354) ce ct +6," or [163] ep eae (P,') = | Po" Ps! Pot +P! P;! (5.35b) P,! P,* P; with Pt = 1/c%. (5.35e) i 4 5-27 The dielectric between the armours (skid wires) and the pipe is represented by the second term in Eq. (5.34). Each of the submatrices Tyg) and [Py] in the second term 4s a 3x3 matrix with 9 equal elements, pon @ 44)? Pac!" tee ange b-(@) (5-368) ryt A oy Gs) a my} oy a, oe 24; dcos0,, n=l a with the essential terms in Eq. (5.36) being the same expressions appearing in Eq. (5.31) and (5.33). The admittance matrix is then found by inverting @ cond! 1 IY cong! Jel" (8.37) cond! cond a 2 At lower frequencies, there {s mutual coupling between the inner and outer surface of the pipe. The induced voltage on the outer eurface will then produce @ circulating current through the pipe and earth return. This extra loop must Le added to the loop impedance matrix of Eq. (5-18), as in Bg. (5.18), with elements defined in Section 5.4.1, except that Eq. (5.32 o 2" oop) | for intinite pipe 4 (6.38a) thickness must be replaced |_0, by a formula fog, finite a pipe thickness ™ 0 ° 2: 00 -2' 00-2" 2 2 2 with uae 2 tpe-mitual from Eq. (5.7c) , (5-38b) a 87 - pipe-out ee insulation * Zearth : (5-38) The first two terms in Eq. (5.38c) are found from Eq. (5-7b) and (5.6) = 0 4f pipe in contact with earth), and 2" is the @ insulation earth ore appears that the support routine CABLE CONSTANTS uses Eq, (5.7a) of the concentric configuration for the non-concentric configuration as well. 5-28 earth-return impedance discussed in Section 5.3. Transforming Eq. (5-38a) to conductor quantities produces same matrix as for infinite (5-39a) 12" cona) pipe thickness OO... with Z,' from Bq. (5.38¢) zt 92,'- 2," (5-39) Bie zt - te ‘The last row and column in Eq. (5-39) represent the pipe quantities, while the first 9 rows and columns refer to core, sheath (shield tapes), armour (skid wires) of phases a, b, and c. If the pipe is in contact with the earth, then the shunt admittance matrix is the same as in Section 5.4.1. If it is insulated, then the potential coefficient matrix of Eq. (5.34) mst be expanded with one extra row and column for the pipe, and the same element * 1 pipe-insulation Ble jn_-Piper insulation (5-40) THESE, “pipe-outside must be added to this expanded matrix, same as in rd Eq. (5-34) G4) The admittance matrix is then again found by inversion with Eq. (5-37). 5.5 Bundling of Conductors and Elimination of Grounded Conductors Conductors are sometimes connected together ("bundled"). For example, the concentric neutral conductors in the cable of Fig. 5.2 are in contact with each other, and therefore electrically connected. In a pipe-type cable, the shield tapes and skid wires are in contact with the pipe. In a submarine 5-29 cable, the sheath is often bonded to the armour at certain intervals, to avoid voltage differences between the sheath and armour. In such cases, the connected conductors 1, m can be replaced by (or bundled into) one equivalent conductor, by introducing the bundling conditions Itt te, vy" 7 fa" Tequiv? 7% Avie, (5-42) fa ~ Yequiv into the equations for the series impedance and shunt admittance matrices i The bundling procedure for reducing the equations from m individual to one equivalent conductor 1s the same as Method 1 of Section 4.1-2-2 for overhead lines, and is therefore not explained again. It is exact if the conductors are continuously connected with zero connection resistance (as the neutral conductors in Fig. 5-2), and accurate enough if the connections are made at discrete points with negligible resistance (as in bonding of the sheath to the armour), as long as the distance between the connection points is short compared to the wavelength of the highest frequency in the transient simulation. As in the case of overhead lines with ground wires, some conductors in a cable may be grounded. For example, the steel pipe of a pipe-type cable can usually be assumed grounded, because its asphalt mastic coating is not an electric insulation. Also, neutral conductors may be connected to ground at certain intervals, or at both ends. If a conductor 1 is grounded, then the condition is simply y70 (5.43) and conductor { can then be eliminated from the system of equations in the same way as described in Section 4.1.2.1. Again, the elimination is only exact 1f the conductor 1s grounded continuously with zero grounding resistance, and accurate enough if the distance between discrete grounding points 1s short compared to the wavelength of the highest frequency. 4n example of bundled as well as grounded conductors would be a single-core submarine cable which has its sheath bonded to the armour. Since the asphalt coating of the armour is not an electric insulation, the armour is in effect in contact with the sea water, and both sheath and armour are therefore grounded conductors. By eliminating both of them, the submarine cable can be represented by single-phase equations for the core conductor, 5-30 with the current return combined in sea water, armour and sheath. For an overhead line, the equivalent situation would be a single-phase line with two ground wires. ‘The case of segmented ground wires in overhead lines discussed in Section 4.1.2.5(b) can exist in cables as well. For example, if the sheath is grounded at one end, but open and ungrounded at the other end, then the sheath could be eliminated in the sane vay as segmented ground wires, provided the cable length is short compared to the wavelength of the highest frequency. The support routine CABLE CONSTANTS does not have an option for such eliminations. The user mist represent the sheath as an explicit conductor, instead, with one end connected to ground. This offers the advantage that the induced voltage at the other end can automatically be obtained, if so desired. 5.6 Buried Pipelines Pipelines buried close to power lines can be subjected to hazardous induction effects, especially during single-line-to-ground faults. To study these effects, one can include the pipeline as an additional conductor into the transmission line representation (Fig. 5.14(a)). For steady-state (2 wi _ Ninduces 2 x (a) polyphase representation (b) single-phase representation Fig. 5.14 ~ Pipeline representation (g = ground wire, a,b,c, = phase conductors, p = pipeline) 5-31 analysis, one can also use the single-phase representation of Fig. 5.14(b), with an impressed voltage ev, induced _ 5, : ‘ : ~ Bed = 2h tat Boply t+ pole + 2 pple: There is no capacitive coupling between the power line and the pipeline if + (5.48) it is buried in the ground. As explained later, nominal n-circuits can only be used for very short lengths of pipeline (typically < 0.3 km at 60 Hz), The single-phase representation is therefore preferable for steady-state analysis, because the distributed parameters of Fig. 5.14(b) are more easily converted into an exact equivalent n-circuit than the polyphase parameters of Fig. 5.14(a). This results in the active equivalent m-circuit of Fig. 5.15, with Y, and Y, series ‘shunt being the usual parameters obtained from Eq. (1.14), while Tinduced y, series Fig, 5.15 - Active equivalent m-circuit is an active current [158], x, induced’ induced ~ 2 juce a The correctness of the active circuit can easily be shown. Starting from the differential equations av ~ ax 7 ppl t 2 pplinducea * Finduced (5.45) ar -dey ax 7 Y'ppY 5-32 the introduction of a modified current T* Taduced Thodified = transforms the differential equations into the normal form of the ine equations, with the assumption that Linduced does not change along the line (dT gai tiea/ a = dI/dx), avon de” 7 pptnodified Ecawotitied oy & pp’ The solution for a line between nodes k and m ie given in Eq. (1-13), except that the current {s now Tse aqs OF rewritten, + GARDy, y, en * “induced series ‘shunt ~ ‘series y, + (/2)¥, y, ak ~ Tinduced| ~ Yoertes Yoeries shunt] | “nl This is exactly the same equation which comes out of the equivalent circuit of Fig. 5.15. With this single-phase approach, the currents in the power line are assumed to be known, e-g-, from the usual type of short-circuit study. It is also assumed that they are constant over the length of the exposure to the pipeline, and that the pipeline runs parallel to the power line (mutual impedances constant). If either assumption is not true, then the power line-pipeline system must be split up into shorter sections as is customarily done in interference studies. The effect of the pipe on the power line zero sequence impedance is usually ignored, but could be taken into account. In both representations of Fig. 5-14, the mtual impedances between the pipe and the overhead conductors, as well as the self impedance of the pipe with earth return, are needed. The mtual impedances are obtained with the formulas discussed in Section 5.3.4. At 60 Hz, Carson's formula will give practically identical results as the more complicated formula of Pollaczek. The self impedance Z',,, of the pipeline consists of the sane three terms shown for the armour in Eq. (5.4). ‘The first two terms are calculated with 5-33 Eq. (5-7b) and (5.6), while R" Section 5-3. For the shunt admittance Y',, = G' + juC', the capacitive part ts arth 18 found from the equations discussed in calculated in the usual way with Eq. (5.13). In contrast to the underground cable, the shunt conductance G' of the pipeline can no longer be ignored. The insulation around pipelines is electrically poor, either originally or because of puncturing during the laying operation. The loss angle 6 in Eq. (5.14) is so large on pipelines insulated with glass-fiber/bitumen that G’ becomes much larger than uC'at power frequency, and if one part of the shunt admittance is ignored it should be uC" rather than G'. On PVC~insulated pipelines, G’ may still be smaller that uC’, though. If the shunt resistance of the insulation ts relatively small, then the grounding resistance of the pipe should be connected in series with ar”) {270}, oF Gis : (5.46) insulation * ¥ grounding where R's tetton Feststance of pipe insulation, R ‘grounding = grounding resistance. A useful formula for the grounding resistance is [170] °, Jon? +h +4 ' ‘earth 22 2? z, oe eee ee (3.47) / z (any? + By - with poarch 7 earth resistivity (e.g-, in Qn), bh = depth of burial of pipe, 4 = length of pipe, D = outside diameter of pipe. * ) If the sheath, armour, or pipe of an underground cable or the ground wire of an overhead line is grounded, then it has been standard practice to ignore the grounding resistance (V='0). An alternative would be to use a finite shunt aduittance Y' = 1/R' sounding? @& Tecently suggested [186]. 5-34 Grounding grids must generally be analyzed as three-dimensional problems, even if they consist of only one pipe. The grounding resistance from Eq. (5-47) 1s therefore no longer an evenly distributed parameter, but depends on the length. Fortunately, the dependence of G' on length is very small for typical values of G'y. oy seyoq [158] In the region of measured values for G' between 0.1 S/km for newly-layed pipelines and 0.3 S/km for older pipelines with glass ftber/bitunen insulation [170], the dependence of G' on length is practically negligible, as shown in Fig. 5.16. Treating G' as an evenly distributed parameter is therefore a reasonable approximation. 10.0 (S/km) Wee Binsulation” ao t oe insulation” 1.0 ns aa tg Ri gnsulacion™ 1000 am 10 100 1000 10000 Fig. 5-16 - Shunt conductance of buried pipe Because of G' >> uC', the wavelength of buried pipelines is significantly shorter than that of underground cables, as shown in Table 5.3 [170]. Therefore, a nominal r-circuit of a circuit which includes a buried 5-35 Table 5.3 - Wavelength of pipeline at 50 Hz [170] G'(S/km) | wavelength (km) 0.1 41.3 1.0 13.1 10.0 4.13 pipeline should not be longer than approximately 0.8 km for steady-state analysis, or approximately 0.08 km for switching surge studies [158]. Fig. 5.17 shows a comparison between measured and calculated voltages and currents in a pipeline, induced by currents in a neighbouring power line, with the pipeline representation as discussed here [158]. Tyipe!Tpouer (PY) 7 be oo oe a JY / - Z Tae 8 eee Se pipe node numbers distance along —— pipeline PIPE VOLTAGE PIPE CURRENT @ BUcker measured 4 BUcker measured © calculated from single-phase @ calculated from single-phase a-model model Tpower ~ fault current in power line Fig. 5.17 - Induced voltages and currents in a buried pipeline 5-36 5.7 Partial Conductor and Finite Element Methods The support routine CABLE CONSTANTS uses analytical formulas which are essentially only applicable to configurations with axial symmetry. The formulas for the nonconcentric configuration in pipe-type cables (Section 5.5) are only approximate, and the authors of these formulas thenselves suggest improvements along the lines discussed here. To find the impedances and capacitances for conductor systems with arbitrary shapes (e-g-, for the cable of Fig. 5.1), numerical methods can be used in place of analytical formulas, which are either based on subdivisions into partial conductors or on finite element methods. There is no support routine yet in the ENTP which uses these nunerical methods. The principle of these methods is therefore only outlined very briefly. 5.7.1 Sut 1 Conduct: With this method, each conductor is subdivided into small “partial” conductors (“subconductors” in [162], “segments” in [164]), as shown in Fig. 5.18. Various shapes can be used for the partial conductors, with rectangles circle square elemental Fig. 5.18 - Subdivision of the main conductors into partial conductors being the preferred shape for strip lines in integrated circuits (Fig. 5-19). 5-37 “ y fasta, (Z , (mapnsynay Z jo2y]r3) [nfomafinay] mes, xe won eH oa on (0) (03), Fig. 5.19 - Subdivision of strip lines into partial conductors of rectangular shape [164]. Copyright 1979 by International Business Machines Corporation; reprinted by permission In deriving the equations for the system of partial conductors, uniform current density {s assumed within each partial conductor. Then the voltage drops along a system of n partial conductors at one frequency are described by the phasor equations av, /ax Ry ak a1 Taz an] | [4] av fax fee ee ee) I 2 Pn) -|. . : + ju] 7 7 - 271 (5.48) av /ax | hay Bap * - * Mantl | Tal ‘The diagonal resistance matrix contains the de resistances, and the full inductance matrix contains the self and mtual inductances of the partial conductors. The formulas for the matrix elements depend on the shape of the partial conductor, but they are well known. To obtain the frequency-dependent impedance of a cable system, the matrices [R] and [L] are first computed. At each frequency, the complex matrix [Z] = [R] + Ju[L] is formed, and reduced to the number of actual conductors with Bundling Method 1 of Section 5.5. For example, if partial conductors 1, «+. 50 belong to the core conductor, and partial conductors 51, + 120 to the sheath, then this bundling procedure will reduce the 120 x 120-matrix to a 2 x 2-matrix, which produces the frequency-dependent 5-38 impedances Zg6M) ZC) This numerical method works well as long as the conductors are subdivided into sufficiently small partial conductors. The size of these partial conductors mist be of the same order of magnitude as the depth of penetration. 5.7.2 Finite Element Methods Finite element methods are more powerful than partial conductor methods in one sense, inasmuch as it is not necessary to assume uniform current density within each element. However, it is very difficult to handle open-boundary conditions with finite element methods, that is configurations where the magnetic field diminishes gradually as one moves away from the conductors, with no clearly defined boundary of known magnetic vector potential reasonably close to the conductors. In situations where a boundary is clearly defined, e-g-, in pipe-type cables at high frequency where the depth of penetration becomes mich less than the wall thickness, finite element methods can be quite useful. With finite element methods, the region inside and outside of the conductors is subdivided into small elements, usually of triangular shape. Fig. 5.20(a) shows the example of a stranded conductor inside a pipe of radius R, as the return path (clearly defined boundary with zero magnetic feild A= 0 outstde the pipe and zero derivative along the to edges of the wedge"). Because of axial symmetry, it is sufficient to analyze the “wedge” shown in Fig. 5.20(a). This wedge region is then subdivided into triangular elements as shown in Fig. 5.20(b), with longer triangles as one moves away from the conductor. ‘The magnetic vector potential A is assumed to vary linearly along the edges and inside of each triangle, Azaxtbyte, 6.49) 5-39 stranded Conductor (b) Subdivision of region into triangular elements Fig. 5-20 - Analysis of stranded conductor with finite element method [171]. Reprinted by permission of Yin Yanan 5-40 when a first-order method is used (higher-order methods exist as well). The unknowns are essentially the values of A in the node points. If they were shown in the z-direction of a three- dimensional picture, then the triangles would appear in a shape similar to a geodesic dome, with the roof height being the value of A. The equations for finding A are linear algebraic equations with a sparse matrix, but the number of node points or the number of equations is usually quite high (146 equations for the example of Fig. 5.20). Once the magnetic vector potential is known in the entire region, the impedances can be derived from it. For readers interested in finite element methods for cable impedance calculations, the Papers by Konrad, Weiss and Csendes [165, 166, 167] are a good introduction. 5.8 Modal Parameters Once the series impedance and shunt admittance matrices per unit length * (2'pnasel: [¥'pnase] are known, the derivation of modal parameters is exactly thé Same as described in Section 4.1.5 for overhead lines. They could be used, for example, to develop exact equivalent m-circuits for steady-state solutions as explained in Section 4.2.1.3. For transient simulations, it is more difficult to use modal parameters, as compared to overhead lines, because the transformation matrix [T,] can no longer be assumed to be constant as for a single-circuit’ overhead line. Fig. 5.21 shows the variation of the elements in the third column of [T,] for a typical three-phase arrangement of 230 kV single-core cables with core conductor and sheath in each [155]. Especially around the Power frequency of 50 or 60 Hz, the variations are quite Pronounced. Above a few kHz, the loop between core conductor and sheath becomes decoupled from the outer loop between sheath and earth return, because the depth of penetration on the inside of the sheath for loop 1 becomes much smaller than the sheath thickness. In that case, Zeype-mutual® 0- This makes the transformation matrix constant above a few kHz, as evident from Fig. 5.21. For a single- phase single-core cable with sheath and armour, the Replacing subscript "cond" of Eq. (5.19) and (5.21), or of similar equations, with "phase" indicates that conductors connected together have been bundled, and grounded conductors have been eliminated (Section 5.5). For example, a three-phase pipe-type cable with its sheaths in contact with the pipe has 7 conductors and [2'. is therefore a 7x7 matrix. [2',~.] is a 4x4 matrix for phase conductors a,b,c, and for the equivalent conductors representing the three sheaths and the pipe bundled together. If the pipe is continuously grounded, it becomes a 3x3 matrix. 5-41 Magnitude of Eleaents of Eigenvector 3 10M Frequency in Hz Fig. 5.21 - Magnitude of the elements of column 3 of [T,] three modes are identical with the 3 loops described in Eq. (5.1) at high frequency where Z',, ~ 0 and Z',, = 0. The transformation matrix (T,] at high frequency is then simply the transformation matrix between loop and conductor quantities of Eq.(5.9), emf t Juoa-f 28 «30 5.9 Cable Models in the EMTP As of now (Summer 1986), there are no specific cable models in the BPA EMTP. The only way to simulate cables is to fit cable data into the models available for overhead lines. It has long been recognized, of course, that this is only possible in a limited number of c: A method specifically developed for cables, as discussed in Section 5.9.2.3, will hopefully be implemented in late 1986 or early 1987. It has already been tested extensively in the UBC EMTP. 5-42 5.9.1. In principle, there is no difficulty in representing cables as nominal or equivalent w-circuits in the same way as overhead lines (Section 4.2.1). If nominal n-circuits are used, it should be realized that the wavelength of underground cables is shorter than on overhead lines. If a nominal x-circuit should not be longer than 100 km at 60 Hz for overhead lines, the limit is more typically 30 km for underground cables. If a pipeline is modelled, the limit can be as low as 1 km, as discussed in Section 5.6. Underground cables are often very short compared to the Length of connected overhead lines. In such cases, the (complicated) series impedances have very little effect on the results because the system sees the cable essentially as a shunt capacitance. The cable can then be modelled as a simple lumped capacitance. 5.9.2 Transient Solutions The accurate representation of cables with frequency-dependent impedances and frequency-dependent transformation matrices is discussed in Section 5.9.2.3. Situations where simpler models should be accurate enough are discussed first. 5.9.2.1 Short Cables If a rectangular wave pulse travels on an overhead line and hits a relatively short underground cable, then the cable termination is essentially seen as a lumped capacitance. The voltage then builds up exponentially with a time constant of T = Z. » as shown in Fig. 5-22(a). If the ‘overhead Ccable’ cable is modelled somewhat more accurately as a lossless distributed-parameter line, then the voltage build-up has the staircase shape of Fig. 5-22(b), with the average of the sending and receiving end curve 5-43 2 overhead cable tie / eo bo (a) Cable represented as lumped (b) Cable represented as capacitance lossless transmission line sending end of cable receiving end of cable Fig. 5.22 - Voltage build-up in a cable connected to an overhead line being more or less the same as the continuous curve in Fig. 5.22(a). As long as the travel time + of the cable is short compared to the time constant 7, reasonably accurate results can be obtained if the cable is represented as a lumped capacitance. Nominal r-circuit representations have often been suggested a approximate cable models. They obviously represent the capacitance effect correctly, but the pronounced frequency-dependence in the series impedance is ignored. Nominal n-circuits give reasonable answers probably only in those cases in which the simpler lumped capacitance representation is already accurate enough. 5.9.2.2 Single-Phase Cables ‘There are situations where single-phase representations are possible. ‘An example is a single-phase submarine cable in which the sheath and armour are bonded together, with the armour being in contact with the sea water. In such a case, the sheath and armour can be eliminated from Eq. (5-10), which 5-4 results in the reduced single-phase equation Mey ae” 7" core*e? with Z' 9, being the impedance of the core conductor with combined current return through sheath, armour and sea water. Coupling to the cables of the other two phases can be ignored because the three cables are layed relatively far apart, to reduce the risk of anchors damaging more than one phase in the same mishap. When the equations have been reduced to single-phase equations, then it is straightforward to use the frequency-dependent overhead line model described in Section 4-2.2.6. Sometimes it is not necessary to take the frequency-dependence in the series impedances into account. For example, single phase SF,-busses have been modelled quite successfully for fast transients with two decoupled lossless single-phase lines, one for the inside coaxial loop and a second one for the outside loop between the enclosure and the earth-return. ‘The coupling between the two loops through the enclosure is negligible at high frequencies because the depth of penetration is much less than the enclosure wall thickness. The only coupling occurs through reflections at the terminations. Agreement between simulation results from such simple models and field tests has been excellent [169]. 5.9.2.3 Polyphase Cables [155] The simple overhead line models with constant parameters discussed in Section 4.2.2 are of limited use for underground cables for two reason: (a) The transformation matrix [T,] is frequency-dependent up to a few Kitz, though a constant [T,] would be acceptable for transients which contain only high frequencies (e-g-, lightning surge studies). (>) The modal parameters (e.g-, wave velocity and attenuation) are more frequency-dependent than on overhead lines, as shown in Pig. 5.23 for three single-core cables with core and sheath [150]. 5-45 attenuation (4B/km) — f (Hz) (a) Attenuation modal velocity (km/ms) © 5 ahs — aces —ea0 — fiz) (b) Velocity Fig. 5.23 - Modal parameters as a function of frequency [150]. Reprinted by permission of TEE and the authors 5-46 To derive an accurate model for an n-conductor cable systen between nodes k and m, we can start from the phasor equation (4.121) for the overhead line, if we replace that scalar equation, which was written for one phase or mode, by a matrix equation for the n conductors, CEI) ~ Ogg) = ALI) + Oy) (3.51) with [¥.1 = (2,17) = characteristic admittance matrix 1n phase quantities, fAl = ett = propagation factor matrix. Eq. (5.51) is transformed to modal quantities, with 11] = IT] tygge) (5-52a) and w= Wagael + (5.52b) which ytelds a-node! * e-node! M-mode! Hnode}{e-mode ! Yn-mode U4 aie-moae! (5-53) with both [¥.aggel and [Ayoge! being diagonal matrices, {Y, = (TL, ney? (5. 54a) ermode! * FATT GMIT T+ om Tooge) 7 ITI IANITy)- (5.54) The diagonal element of [A,.,,] is obtained from the i-th eigenvalue A, of the product 1¥" 450112" agels - BVA Aodet 7 © (5.540) and [T,] 18 the matrix of eigenvectors of the sane product (phase! !2"phase!* #4- (5-53) consists of n decoupled (scalar) equations, with one equation for each mode. Transforming these scalar equations into the time domain is the same procedure as described in Section 4.2.2.6 for the overhead line. For mode 4, the second term in Eq. (5-53) is found with the sane convolution integral as in Eq. (4.124), his! 1 (t ~ w)a(u)du, for each mode (5.55) ‘propagation ~ J tutota: 5-47 with the current 1, .5¢g betmg the sum of the line current iy and of @ current which would flow through the characteristic impedance of mode i if the voltage v, of mode 4 were connected across it. Only known history terms appear in Bq (5-55), and hist, rapqrion cam therefore be found by a y as in Section recursive convolutions for the n modes, in the same 4.2.2.6. The modal propagation factors are very similar in shape to those of an overhead line, as shown for A.,4,_3 (w) in Fig. 5.24. Magnitude of Elenent 3 of AY 10 10 I Frequency in Hz Fig. 5.24 - Magnitude of propagation factor for mode 3 of a 6-conductor system (three single-core cables with core and sheath in each) With propagation of the conditions from m to k being taken care of through Eq. (5.55), the only unresolved issue in the modal domain equations is the representation of the term ¥.V, in Eq. (5-53). Again, the frequency dependence of Y, 1s similar to that of an overhead line, as show in Fig. 5.25, and can be represented with the sane type of Foster-I R-C network shown in Fig. 4.42(a), and reproduced here as Fig. 5.26. By applying the trapezoidal rule of integration to the capacitances, or by using recursive 5-48 os oa 035 030 025 020 os 010 005 Magnitude of Elesent 3 of Yc° (ahos? 10 Frequency in H2 Fig. 5.25 - Magnitude of characteristic admittance for mode 3 (same 6-conductor system as in Fig. 5.24) Bist, opagation \ nist : Rc ‘ ¢ eww > [i] fy Py I a - + 7 ist opagation (a) R-C network (b) Equivalent resistance after vplying iaplicit.tategration tere Fig. 5.26 - Representation of one mode seen from side k 5-49 convolution as discussed in Appendix V, the R-C network is converted into an equivalent conductance Cayty + hist . After the network solution at each time step, the current propagation flowing through the characteristic impedance represented by the R-C network in parallel with a known current source histyo aust be calculated for both ends of the cable from G.,,,,¥ + histgo, because this term is needed after the elapse of travel time to form the expression tecorar Needed in Bq. (5.55). From Fig. 5-26(b) it can be seen that each mode is now represented by the scalar, algebraic equation tyg(t) = Squiyt(t) # (hiistys + MSE gation) .56) with an analogous equation for i) (t) at the other end. If the transformation matrix were constant and real, then Eq. (5.56) could very easily be transformed back to phase quantities, (e)] = (T116, J + [Ty] [hist] t Uya-phase ‘equiv! 74] [4 -phase’ as explained in Eq. (4-109) for the overhead line. As shown in Fig. 5.21, the transformation matrix [T,] of cables is very mich frequency-dependent, and the transformation back to phase quantities now requires convolutions based on Eq- (5-52), t Cols f fey @-wil, t Wapae6t)) = J [eC] "Ey age) 124 (5.57) (a) ]du,, (5.57a) Uphase ‘mode! where [t,] 1s a matrix obtained from the inverse Fourier transform of the frequency-dependent matrix [T,]. Similar to the curve fitting used for the modal characteristic impedances, each element of [T,] is again approximated by rational functions of the form ky 7, Co) = ky “i wy (5-58) 5-50 with k,, k, and p, being real constants which, when transformed into the time domain, becomes tyyCED 9 HgBCED +, genm(-pze)ace) (5.59) With the simple sum of exponentials in Eq. (5.59), recursive convolution can be applied again (Appendix V). Then, the convolution integrals in Eq. (5.57) can be split up into a term containing the yet unknown voltages and currents at time t, and the known history terms which can be updated recursively, CE] = [tg}tagge(t] + Ihe erent)» (5.60a) (e)] + [hist, 1, (5.600) i, phase Wade (t)] = [tl phase’ with [t,] being a real, constant n x n-matrix. With Eq. (5.60), the voltage transformation of the modal equations (5.56) to phase quantities is now fairly simple, Upgephase ®t?) * (Spnage]-phase6t?] + (Mt pnage] > (S.61a) with t (nage) 7 (ol ISphasel ty] * > (5.610) and the history term Uhistynase! ~ (M4 Steurrent! + [to] (Coquivl IM styortage! + [histyo] + (hist, -opagation!! (G-6le) Since the form of Eq. (5.6la) is identical to that of Eq. (4.109) for the overhead line with constant [T,], adding the model to the EMIP is the same as described there. The extra effort lies essentially in the evaluation current) 24 [histyoipage]* After the network solution at each time step, Eq. (5.60) is used to obtain the modal of the two extra history vectors [hist, quantities from the phase quantities. ‘The principle of the frequency-dependent cable model is fairly simple, but its correct implenentation depends on many intricacies, which are described in [155]. In particular, it is important to normalize the eigenvectors in such a way that the elements of [T,] as well as the modal surge admittances ¥ both become minimum phase shift functions. This ‘e~mode-i is achieved by making one element of each eigenvector a real and constant 5-51 number through the entire frequency range. Furthermore, standard eigenvalue/eigenvector subroutines do not produce smooth curves of (T,] and Wy, ‘c~mode eigenvalues are calculated often changes as one moves from one frequency ] as a function of frequency, because the order in which the point to the next. This problem was solved by using an extension of the Jacobi method for complex symmetric matrices. Symmetry is obtained by reformulating the eigenproblen 1 ase! (2 "phase! X= Mx] in the form ()(e] = Ate}, (5.62a) where (H) = (L]*[2"phase} [L] , (5-62) and tx} = GE), (5.620) with [L] being the lower triangular matrix obtained from the Choleski decomposition of [¥" nage] (157]+ The Choleski decomposition is a modification of the Gauss elimination method, as explained in Appendix III. One can also replace {L] in Eq. (5-62) with the square root of [¥" 4.40] phase obtained from 1.2 2 (phase! 7 ILA 1X (5-63) where [A1/2] is the diagonal matrix of the square roots of the eigenvalues, and [X] ie the eigenvector matrix of [Y" 1. Both approaches are very phase efficient if G! is ignored, or if tan6 is constant for all dielectrics in the cable system, because [L] or [¥" ]}/2 gust then only be computed once for phase all frequencies. For parallel single core cables layed in the ground (not in air), [¥"] is diagonal if loop equations are used. In that case it is more efficient to oop) 2"zoopl+ where both [L] and [¥", op] 1/2 become the same diagonal matrix with /Y™)o.0 yas its elements. The conversion back to phase quantities is trivial with Eq. find the eigenvalues and eigenvectors for [¥" (5.50) if that equation is expanded from single-phase to more phases. 5-92 The reason why the Jacobi procedure produces smooth eigenvectors is that the Jacobi algorithm requires an initial guess for the solution of the eigenvectors. This initial guess is readily available from the solution of the eigenproblem of the preceding frequency step; consequently, the order of the eigenvectors from one calculation to the next is not lost. Figure 5.27(a) shows the magnitude of the elements of row 3 of the eigenvector matrix (T,] for the same 6-conductor system as in Fig. 5-24, when standard eigenvalue/eigenvector routines are used. Fig. 5.27(b), on the other hand, shows the same elements of [T,] calculated with the modified Jacobi algorithm. As an application for this cable model, consider the case of three 230 kV single-core cables (with core and sheath), buried side by side in horizontal configuration, with a length of 10 km. A unit-step voltage is applied to the core of phase A, and the cores of phases B and C as well as all three sheaths are left ungrounded at both ends. The unit-step function was approximated as a periodic rectangular pulse of 10 ms duration and a period of 20 ms with a Fourier series containing 500 harmonics, 500 WED = as +) CageoeCaye) + bystntuye)) - The wave front of this approximation is shown in Fig. 5.28. Choosing a 10 00 0 0 0 60 30 «0 30 20 10 -0.00 4 0.10 Step Response: Input voltage (ou) 0 05 10 1:5 20 25 3.0 35 40 4.5 5.0 Time (Milliseconds) Fig. 5-28 - Fourier series approximation of unit-step 5-53 & g 2.0 Ee Bs ei Wane? 6 me Zoe B08 yoe = 02 Foo 10 I Ml M 0" ReOUENCY HZ (a) Standard eigenvalue/eigenvector subroutines to og 08 USING JACOBI MAGNITUDE OF ROW 3 OF TT Te bee ee Lem VO ed gee 90 ees 10-10 ee 10) FREQUENCY IN H2 (b) Modified Jacobi algorithm Fig. 5.27 - Magnitude of the elements of row 3 of (T,] (same 6-conductor system as in Fig. 5.24) 5-54 Fourier series approximation for the voltage source offered the advantage that exact answers could be found as well, by using ac steady-state solutions with exact equivalent n-circuits (Section 4.2.1.3) at each of the 500 frequencies, and by euperimposing then. Fig. 5.29 and 5.30 show the EMTP simulation results in the region of the third pulse, superimposed on the exact answers. The two curves are indistinguishable in this third pulse region where the phenonena have already becone more or less periodic. This shows that the EMTP cable model is capable of producing highly accurate answers. The insert on the right-hand side of Fig. 5.29 shows the response to the first pulse, where the EMTP simulation results differ slightly from the exact answers, not because of inaccuracies in the model but because the ENTP starts from zero initial conditions while the exact answer assumes periodic behaviour even for t < 0. 1.50 1.20 0.90 ‘ore 1 Voltage (pu) 0.60 0.30 -0.00 0.30 eet a 4 4 43. 44. 45 46 a7? SO 2 4 Tige (i Lliseconds) Fig. 5-29 - Step response, receiving end voltage of core (phase A) 5-55 0.50 0.40 0.30 0.20 Step Response: Sheath 1 Voltage (pu) 4) a 43 4a 45 Ga Tise (tuLlisecands) Fig. 5.30 - Step response, receiving end voltage of sheath (phase A) 6-1 6. ‘TRANSFORMERS The first representation of transformers in the EMTP was in the form of branch resistance and inductance matrices [R] and [1]. The support routine XFORMER was written to produce these matrices from the test data of single-phase two- and three-winding transformers. Stray capacitances are ignored in these representations, and they are therefore only valid up to a few kiiz. A star circuit representation for N-winding transformers (called “saturable transformer component” in the BPA EMTP) was added later, which uses matrices [R] and [1]~! with the alternate equation (1]> [v] = [2]- [Rr] [4] + [at/ae) (6-1) in the transient solution. This formulation also became useful when support routines BCTRAN and TRELEG were developed for inductance and inverse inductance matrix representations of three-phase units. An attempt was made to extend the star circuit to three-phase units as well, through the addition of a zero-sequence air-return path reluctance. This model has seldom been used, however, because the zero-sequence reluctance value is difficult to obtain. Saturation effects have been modelled by adding extra nonlinear inductance and resistance branches to the inductance or inverse inductance matrix representations, or in the case of the star circuit, with the built-in nonlinear magnetizing inductance and iron-core resistance. A nonlinear inductance with hysteresis effects (called “pseudo-nonlinear hysteretic reactor” in the BPA EMTP) has been developed as well. An accurate representation of hysteresis and eddy current effects, of skin effect in the coils, and of stray capacitance effects is still difficult at this time, and some progress in modelling these effects can be expected in the years to Surprisingly, the simplest transformer representation in the form of an “tdeal” transformer was the last model to be added to the EMTP in 1982, as part of a revision to allow for voltage sources between nodes. 6-2 6.1 Transformers as Part of Thevenin Equivalent Circuits If a disturbance occurs on the high side of a step-up transformer, then the network behind that transformer, plus the transformer itself, is usually represented as a voltage source behind R-L branches. Since the transformer inductances tend to filter out the high frequencies, such a low-frequency R-L cireuit appears to be reasonable. To explain the derivation of such Thevenin equivalent circuits, the practical example of Fig. 6-1 shall be used [80], where the feeding network consists of three generators and two three-winding transformers. The transformer short-circuit reactances ate Xy) = O+117 puss Xyp = Oe115 potey RECEIVING 345KV-1ine EXD (open) 398kn three-winding transformers The 4th generator was disconnected for acceptance testing, Fig. 6.1 - Network configuration for various field tests at CEMIG, Brazil [80] X,q = 0-261 peur, and the generator reactance is Xj = 0.1385 p.u., all based on 100 MVA at 60 Hz. With the well-known equivalent star circuit for three-winding transformers (see Section 6.3.2), the power plant in Fig. 6-1 can be represented with the positive and zero sequence networks of Fig. 6.2. For simplicity, resistances are ignored, but they could easily be included. It 4s further assumed here that the zero sequence reactance values of the 6-3 transforser are the same as the positive sequence values, which is only correct for three-phase banks built from single phase units, but not quite correct for three-phase units (1f the zero sequence values were known, then 0.1385 H H f H H H 0.1195 (a) Positive sequence (negative () Zero sequence sequence identical, except that voltage sources are shorted) Fig. 6.2 - Equivalent circuits for the power plant (reactance values in p-u. based on 100 NVA at 60 Hz) those values could of course be used in Fig. 6.2(b)). Furthermore, the generator is modelled as a symmetrical voltage source E” behind X3. Note that the delta-connected windings act as short-circufts for zero sequence currents in Fig. 6-2(b), while the generators are disconnected to force Tugeo 7 0+ ‘The zero sequence paraneters of the generators are therefore irrelevant in thie example. The networks of Fig. 6-2 can now be reduced to the three Thevenin equivalent circuits of Fig. 6.3, which in turn can be converted to one three-phase Thevenin equivalent circuit as shown in Fig. 6.4. This three-phase circuit is used in the EMTP for the representation of the power plant, with the data usually converted from p-u. to actual values seen from the 345 KV side (Ko, = Xugg = 99-90 & Kop. 7 3917 @ oF X, = 77.65 0 X, 6-4 = -22.25 Qat 60 Hz). The symmetrical voltage sources Ey Ey FE, behind the coupled inductances in Fig. 6.4 are the open-cireult voltages of the pover plont on the 345 KV side. In the transient simulation, the matrix [x] ts obviously replaced by the inductance matrix [L]- 0.08393 p.u. 0.08393 p.u. 0.02787 p.u. pos. seq. neg. seq. zero seq. Fig. 6.3 ~ Thevenin equivalent circuits in sequence quantities XX | ~ coupled reactances ee xe Xp Xa Xs X, = 0.06524 peur } at 60 He Xq = -0.01869 peu. Fig. 6.4 - Three-phase Thevenin equivalent circuit in phase quantities 6.2 Inductance Matrix Representation of Single-Phase Two- and Three-Winding Transformers Transformers can only be represented as coupled [R}{L branches 1f the exciting current is not ignored. The derivations are fairly simple, and shall be explained with specific examples. 6.2.1 Ivo-Winding Transformers Aseume a short-circuit reactance of 10%, short-circuit losses of 0.5%, 6-5 and an exciting current of 1%, based on the ratings V, of the rating? Srating transformer. The excitation losses are ignored, but could be taken into account as explained in Section 6.6. If the given quantities are Z,,, load losses P,,.45 and power rating S, » then the resistance and reactance rating part of the short-circuit impedance are Rou” Pross!/Srating (6.28) x2 S22, = R2 (6.20) pu pu” “pu Since the load losses do not give any information about their distribution between windings 1 and 2, it is best to assume If the winding resistances are known, and not calculated from P,,.., then Ry pu amd Ry 4, may of course be different, and R= 8, ), +R, ,, 18 then used in Eq. (6-2b). With the T-circult representation found in most textbooks, the p-u. impedances are then as shown in Fig. 6.5. The short-circuit impedance 0-005 + 0.10 p-u. is divided into two equal parts, and the magnetizing reactance 499.95 p-u., which 1s purely imaginary when excitation losses are ignored, is chosen to give an input impedance of 100 Psu. from one side, with the other side open, to make the exciting current 0.01 peu. (the resistance 0.0025 p.u. 1s so small compared to 100 p-u. that it can be ignored in finding the value 99-95). The equations with the branch impedance matrix in p.u. are then 0,00254j0.05 p.u, 0.0025+j0.05 p.u. ® @ 599.95 peu. Fig. 6.5 - T-circuft representation of transformer vy, 1 pu | _ | [70.0028 =o +57 200 99.95 wu] 63a) v, ° 0.0025, 99.95 100 _| Topu for steady-state solutions, or Vy] = (ey) [42] + (yy | Atal (6-3b) v 1 ai ,/at for transient solutions, with [R] being the same matrix as in Eq. (6.3a), and (L] =4 [x]. Most MTP studies are done with actual values rather then with peu. values. In that case, the matrix in Eq- (6-3) must be converted to actual values, with 0.0025 va 100 ve 99.95, Vio (l-3 @ (6.4) Seating © 0.0025 v3 99.95 v,v, 100 v2 2 where S - er rating of transformer, © Seating ~ SPParent power rating of transformer Ve Vy = voltage ratings of transformer. Eq+ (6-4) gives the (R] and [X]-matrices of coupled branches in Q, as required by the EMTP, with the correct turns ratio Vy/Nys If all quantities are to be referred to one side, say side 1, then simply set V) = V, in Bq. (664). It 4s 4mportant to realize that the branch impedance matrix [2] tn Bq. (6.4) does not imply that the two coupled branches be connected as shown in the Tcireuit of Fig. 6.5. If it were indeed limited to that connection, one could not represent a three-phase bank in wye/delta connection, because both sides would alvays be connected from node to ground or to some other comaon node. Instead, [2] simply represents two coupled coils (Fig. 6.6). The connections are only defined through node name assigmments. For example, if three single-phase transformers are connected as a three-phase bank with a grounded wye connection on side 1 and a delta connection on side 2, then the first transformer could have its two coupled branches from node HA to ground 6-7 and from LA to LB, the second transformer from HB to ground and LB to LC, and the third transformer from HC to ground and LC to LA. This connection will also create the correct phase shift automatically (side 2 lagging behind side 1 by 30° for balanced positive sequence operation in this particular case). 22 Fig. 6.6 - Two coupled coils the four elements in the [X}oatrix of Eq. (6+3) contain basically the information for the exciting current (magnetizing reactance X, = 100 peu.), with the short-circuit reactance being represented indirectly through the snall differences between X,1 and X1>, and between Xz and Xp1+ Tf all four values were rounded to one digit behind the decimal point (X1, = Xp. ~ X12 = 100 pou), then the short circuit reactance would be conpletely lost cxShort = 0). In most studies, it is the short-circuit reactance rather than the magnetizing reactance, however, which influences the results. It is therefore important that [X] be calculated and put into the data file with very high accuracy (typically with at least 5 or 6 digits), to make certain that the short-efreuft reactance 2 X12 short . x1) - = » seen from side 1, (6.5) Xa2 fh short is still reasonably accurate. It is highly recommended to calculate X' from Eq. (645), to check how much it differs from the original test data. For a transformer with 10% short-circuit reactance and 0.4% exciting current, the values of Z,)) 212» Zz2 would have to be accurate to within 0.001% to 6-8 achieve an accuracy of #10% for x°"°¥) this accuracy problem is one of the reasons why Z,,, 2,7, Zp) cannot be measured directly in tests if this data is to contain the short-circuit test information besides the excitation test information. Mathematically, [X| is almost singular and therefore ill-conditioned, the more so the smaller the exciting current is. Experience has shown that the inversion of [X] inside the EMTP does not cause any problens, as long as very high accuracy is used in the input data, Problems nay appear on low-precision computers, however. The author therefore prefers inverse inductance matrix representations, as discussed in Section 6.3. The impedance matrix of single-phase three-winding transformers can be obtained in a similar way with the well-known star circuit used in Fig. 6.2. In that circuit, the magnetizing reactance 1s usually connected to the star point, but since its unsaturated value is much larger than the short-circuit reactances, it could be connected to either the primary, secondary or tertiary side as well. Assuming that the exciting current for the example of Fig. 6.2 4s 1% measured from the primary side, with excitation losses ignored, the magnetizing reactance in the star point would then be 100.0045 peu. Then 100 100.0045 100.004: [x] = | 100.0045 100.1260 100.0045 | p.u. (6.6) 100.0045 100.0045 100.1240 ‘The particular connection would again be established through the node names at both ends of the branches. For example, the three branches could be connected from node HA to ground, LA to LB, and TA to TB. To convert Eq. (6-6) to actual values, divide all elements by the power rating §, > and nuleiply che ftrae rov and coluan with voleage eating Vj, the second row and colum with V,, and the third row and column with Vj- The (R]- and [X]-matrices can either be derived by hand, or they can be obtained from the support routines XFORMER, BCTRAN, or TRELEG in the BPA version of the EMTP. The latter two support routines were developed for 6-9 three-phase units, but can be used for single-phase units as well. 6.3 se _Induct: ix Repres. 1 Single-Phase Two- ind_Three-Wind iy formers If the exciting current is ignored, then the only way to represent transformers is with branch matrices [R] and [L]~2’ which are handled by the EMTP as described in Section 3.4.2. The author Prefers this representation over all others, because the matrices {R] and [L]"? are not ill-conditioned, and’ because any value of exciting current, including zero, can be used. The built-in star circuit in the BPA version of the EMIP uses this representation internally as well. For three~ to [R]-and [L) i with the support routine BCTRAN. For single-phase units and for three-phase transformers Where Zzero * Zgog, the conversion is fairly simple, and can easily be done by hand," as explained next. 6.3.1 wo Winding Transformers First separate the short-circuit impedance into its resistance and reactance part with Eq. (6.2). The [R]- and [wL]~! branch matrices in p.u. can then be written down by inspection from the equivalent circuit of Fig. 6.5 (after the magnetizing inductance has been removed) , 1 Rip 0 and (aL) ‘a (6.7) Oo Rape L Xpu The inverse branch reactance matrix [wL,,)~! leads to the well-known node admittance matrix of a seried“branch with p.u. reactance X,,, if the two bottom nodes in Fig. 6.6 on sides 1 and 2 are grounded. If they are not grounded, then the 2 x 2 branch matrix enters into 4 locations in the formation of the admittance matrix, as explained in Fig. 3.2 (for 3 coupled branches there, compared to 2 coupled branches here) . As shown in Fig. 3.3, two coupled branches can also be represented by 6 uncoupled branches. In programs which do not accept coupled branches this is a useful alternative. In the EMTP, it is more straightforward to use the coupled-branch option. For the example of Fig. 6.5, with exciting current ignored, the p.u. matrices would be 6-10 0.0025 0 ul=| 0 — o.0025 The matrices in Eq. (6.7) are converted to actual values with v2 L Ipu “1 ° [a] = ing (6.98) Seating 2 rating | 0 Roy V2 -2 Seating | Vj? VyV. {wi}? = ating | Yy 2 | ins (6-9) Las 2 1 2 vy, YY With S,aeing ” @PPAFeNE power rating, V1, Vy = voltage ratings. Bq. (6-9) contains the correct turns ratio V,/V,. If all quantities are to be referred to one side, say side 1, then simply set V, = V, in Bq. (6-9)+ To obtain [L]“}, the matrix in Eq. (6.9) fs simply multiplied with u. As already mentioned in Section 3.1.2, the two coupled branches described by Eq. (6.9) can also be represented as six uncoupled branches- Ignoring the resistances for the sake of this argument, and setting 8, y= fetng ou vy t= ‘1 % produces the steady-state branch equations (3.3) and the alternate representations with uncoupled branches of Fig. 3.3. Separating R and X is more complicated now. Therefore, R shall be 6-11 ignored in the following explanations. Resistances can be included, however, 4£ the support routines BCTRAN or TRELEG are used (see Section 6.10.2 and 6.10.3). The starting point is the well-known star circuit of Fig. 6.7. ? Fig. 6.7 - Star circuit for three-winding transformer with p-u. values based on voltage ratings, or with actual values referred to one side Its reactances are found from the pu. short-circuit reactances Xiu» Xyrpy? base’ Since the power transfer ratings Sy) between H-L, S,_ between H-T, and S,_ between Xyrpur based on the voltage ratings and one common power base 5, L-T are usually not identical, a power base conversion is usually needed. If we choose S,,., as the common base (in same units as power ratings Sis Supe Spp)s then x x Sage i Cae « Mitow Bato. g, pu Sent ae Se) * Shase Xie “us % 1 apy , Spy _ “utp Sipu 2 (Cea + Tauae - Ftp). g (6-10) LT a Sar base x, x x, 1 (Sunpu rm Xopu 7 (eee + tee - Faas « Ser Sur aL ‘base For the example used in Section 6.1, with X= 0.117 p.us, Xyp= 0-115 pete, X77 0-241 p.u., these star-circuit reactances would be Xypu™ 70-0045, X= 041215, Xp,,= 0.1195 Next, the well-known star-delta transformation is used to convert the 6-12 star-circuit of Fig. 6.7 into the delta circuit of Fig. 6.8, L Frnpy B mpu impo T Fig. 6.8 - Delta circuit which gives us the susceptances” - Spe Butpu” “ys - Te (6.11a) Bam” 3 3. Tips Ltpu ” ys with © = p04 Xpu + Xtpu Xtpu * Sipy “pu (6.11b) For the numerical example, Brput 889484, Bypyyt 9-437, Bug yt — 0433495. Note that the susceptances in Eq.(6.1la) are not the reciprocals of the ma short-circuit reactances X used in Eq.(6.10). The p-u, matrix [¥t,.,] is easily obtained from Fig. 6.8 with the rules for nodal admittance matrices as *)"susceptance” B is used here for the reciprocal of reactance X. This is not strictly correct, because susceptance is the imaginary part of an admittance (which implies B = -1/X). 6-13 Bunga Bima — ~Bzpu Biya (6.12) fot) =| ~Bapu Bugut Bru ~Brzpu ~Bimpu ~Brtpu Brmput Brap or for the numerical example, [17.93855 -8.89484 -9.04371] [@L,]* = |-8.89484 8.55989 0.33495 -9.04371 0.33495 8.70876. The matrix [wL}"+ in actual values is found as 1st row and column of (6.12) x Vissi ‘a [wz] = |and row and column of (6.12) x Liason in $(6-13) 3rd row and column of (6.12) x Wises a This matrix will contain the correct turns ratios. If all quantities are to be referred to one side, say side H, then simply set V.= Vz= Vy in Eq. (6.13). “Phase N-Coi sformers The newer support routines BCTRAN and TRELEG are not limited to the particular case of two or three coils, but work for any number of coils. If each winding is represented as only one coil *), then transformers with more than three coils will seldom be encountered, but if each winding is represented as an assembly of coils, then transformer models for more than three coils are definitely needed. Breaking one winding up into an assembly of coils may well be required for yet to be developed high-frequency ") A coil is "an assemblage of successive convolutions of a conductor", whereas a winding is "an assembly of coils" [76]. Since a winding may either be represented as one or as more coils, the more general term "coil" is used here. 6-14 models with stray capacitances. To explain the concept, only single-phase N-coil transformers are considered in this section. The extension to three-phase units is described in Section 6.5. For such an N-coil transformer, the steady-state equations with a branch impedance matrix [Z] are 7 741 712 "ay a fap san % ee (6.14) 2m fz 1 The matrix in Eq. (6.14) 1s symmetric. Its elements could theoretically be measured in excitation tests: If coil k is energized, and all other coils are open-cireuited, then the measured values for I, and Vj,---Vy produce colum k of the [2] matrix, 2a 7 Vy/Ty (6.15) Unfortunately, the short-circuit impedances, which describe the more important transfer characteristics of the transformer, get lost in such excitation measurements, as mentioned in Section 6.2. It is therefore much better to use the branch admittance matrix formulation (1) = ([y][v) (6.16) which {s the inverse relationship of Eq. (6-14). Even though [Z] becomes infinite for zero exciting current, or ill-conditioned for very small exciting currents, [¥] does exist, and is in fact the well-known representation of transformers used in power flow studies. Furthermore, all elements of [¥] can be obtained directly from the standard short-circuit test data, without having to use any equivalent circuits. This is especially important for N > 3, because the star-circuit (“saturable transformer component” in the BPA EMTP) is incorrect for more than three coils. For an intermediate step in obtaining [¥], the transfer characteristics between coils are needed. Let these transfer characteristics be expressed as 6-15 voltage drops between coil i and the last cotl N, reduced ,reduced reduced Vy ‘11 212 7 2h we. - - reduced ,reduced reduced | | 1 yy 21 2 cee eae 2 . - 5 (6.17) reduced reduced reduced Mean, eit 7yaj2 tt Zyeaywea_| | Tea, with [2"educed) again being symmetric. Since the exciting current has negligible influence on these transfer the exciting current altogether. Then (based on one common base power §, ‘base? ase of the N coils) must be zero, or N 1 ma * ‘The p.u. values of the matrix elements characteristics, it is best to ignore *) and on the transformer voltage ratings the sum of the p-u. currents =0 pu (6.18) in Eq. (6-17) can then be found directly from the short-circuit test data, as first shown by Shipley [108]. For a short-circuit test between i and nonzero, and Vy), = 0+ Then the i-th N, only Ty py 4m Eq. (6-17) ts row becomes reduced M4 pu 7 44 pot pos (6.19) ‘The impedance in this equation is the short-circuit impedance between coils i and N by definition, reduced _ short tt po tN pa? (6.20) reduced based on one common base power S . The off-diagonal element 24¢°UC°* is base ik pu found by relating rows i and k of Eq. (6-17) to the short-circuit test *)erom here on it is best to work with P+us quantities, or with quantities referred to one side, to avoid carrying the turns ratios through all the derivations. 6-16 between 1 and kK. For ehis test, T 4 = Ty ys and Vy, = 0, with all other currents being zero. Then rows i and k become = (qreduced _ preduced) Ye pu Ywpu ir pu” Atk po) pu * (6-210) reduced _ ,reduced Ya pu "(Ad pe 7 ee pe JL pu? ee reduced | ,reduced or after subtracting Eq. (6-21b) from (6.21a), with 2¢¢ = 2 5 = (gteduced , ,reduced _ , reduced “4 pu 7 Cr pu * Ace pu ik pet pu * (6-210) By definition, the expression in parenthesis of Eq. (6-21c) must be the short short-cireult impedance 21/°C', or 1 (yshort , yshort _ ,short 2 (pe * Aen pu ~ 2K pu! (6.22) based on one common base power S,,.,. This completes the calculation of the base matrix elements of Eq- (6.17) from the short-circuit test data, which is normally supplied by the manufacturer. Eq.(6-17) cannot be expanded to include all coils, since all matrix elements would become infinite with the exciting current being ignored. The impedance matrix of Eq.(6.14) can therefore not be used. To get to the admittance matrix formulation (6.16), Eq.(6-17) is first inverted, [yeduced) jteduced) 1, (6.23) In this inverse relationship, the voltage Vy ,, of the last cofl already exists, and all terns associated with it can be collected {ato a N-th column for Vy pyr The Nth row fe created by taking the negative sum of rows L,+ssN-1 based on Eq- (6-18). This results in the full matrix representation we OY ¥, YL pw Yi pu “12 po IN pu 1 pu 1 x, y on y, 2 pu 21 pu “22 pu 28 pu 2 pu . : 5 » (6.244) an @ \, Ty pu Yw1 pu Yn2 pu NN. pu. IN pu. 6-17 with y, = yreduced ¢ (6-23) for 1, k < Nel 6 240) tke pu” Yin pu fom Fa- (6 or i, (6 +24) Nel reduced Yow pu” “nt pu ais pu for ie (6.24e) Nel Yn pa ~ an pe (6.244) To convert from p.u. to actual values, all elements in Eq. (6-24) are aultiplied by the one comnon base pover S,,,,, and each row and column i is multiplied with VN For transient stuiies, the resistance and inductance parts must be separated, ina way similar to that of Section 6.3. This 1s best accomplished by building [2™°¢%*4] only from the reactance part of the short-circuit test data, which is store _ /;ebort yy ; xghore «J (2tort yay +R 5) (6.25) with short = short=cdi dt it id G Mtud en ashore Breer teeeceeslcernccis)| Ry put Bc pu 7 efther peur load losses tn short-eircuit test between i and k, or sum of p.u. winding resistances. ‘The winding resistances then form a diagonal matrix [R], and 1 GY" = 6M ysenoue r} C with (Yycthout R! being purely built from reactance values jwl. Both [R] and [L]"+ are used in Eq.(6.1) to represent the N-coil transformer. Support routine BCTRAN uses this procedure for obtaining [R] and [L}! from the transformer test data, with two additional refinements: a+ If the winding resistances are not given, but the load losses in the short-circuft tests are known, then the resistances can be calculated from Eq. (6-2) for N=2, and from the following three 6-18 equations for N=3, loss Rw lo: pe + 8s pu 7 P13 pw (6-27) loss R2 pu * ®3 pu” P23 pu Strictly speaking, Eq. (6-2) and (6-27) are not quite correct, because the load losses contain stray losses in addition to the T?R-losses, but the results should be reasonable. For transformers with 4 or more coils there is no easy way to find resistances from the load losses, and coil resistances must be specified as input data 1f N> 4. Additional branches can be added to represent the exciting current, as described in Section 6.6. To show the derivations for a numerical example, let us first use the two-winding transformer of Fig. 6.5, with exciting current ignored. The resistance and reactance part is already separated in this case, with Fyq 7 0,005 ond X,, “0-10: The reduced reactance macrix of Fq- (6.17) 4s pu = §0.10, and its inverse is the = -j10. Adding a second row and column with Eq. (6.24) just a scalar in this case, jx’ reciprocal yreduced pu produces Lf ja et] 2° 7 zloty? =F Lo io] + whch, rogether with Ry y= Rp py = 0.0025, de the sane result show in Ba. (6.8). For the example of the three-vinding transforser used after Eq. (6.10), the reduced reactance matrix (without the factor 4) ts [areduced) -1150 0.1195 ‘pu +1195 0.2410, 6-19 which, after inversion, becomes 17.9386 -8.8948) 8.8948 8.5599 a J or after adding the third row and column with Eq. (6.24), 17.93856 -8.89484 -9,04372 -8,89484 8.55989 0.33495 -9.04372 0.33495 8.70877 [%p_l = ule which is the same answer as the one given after Eq. (6.12), except for minor round-off errors. To convert this matrix to actual units, the rescaling of Eq. (6.13) would have to be used. Since this example from Fig. 6.1 and 6.2 is from a three-phase bank consisting of single-phase units, the ratings to be used are ating = 100/3 MVA, Vy = 345/V¥3 kV (line-to-ground voltage across = nidifig on wye-connected side), V, = Vz = 13.8 kV (line-to-line across winding on delta-connected sides, assuming that 13.8 kV is the rated voltage on the delta side). The star-circuit equivalent circuit of a three-winding transformer is therefore just a special case of the general method for N coils discussed here. 6.5 Matrix tation of Three-Phase N-Coil Transforne: The first attempt to extend single-phase to three-phase transformer models was the addition of a zero-sequence reluctance to the equivalent star-circuit ("saturable transformer element" in the BPA EMTP). This was similar to the approach used on transient network analyzers, where magnetic coupling among the three core legs is usually modelled with the addition of extra delta-connected windings to a three-phase bank consisting of single-phase units. To relate the available test data to the data of the added winding is unfortunately difficult, if not impossible. For example, a two- winding three-phase unit is characterized by only two short-circuit impedances (one from the positive sequence test, and the other from the zero sequence test). Adding delta-connected windings to single-phase two-winding transformers would require three short- circuit impedances, however, because this trick converts the model into a three-winding transformer. adding extra delta-connected windings becomes even more complicated for three-phase three- winding units, not only in fitting the model data to the test data, but also because a four-winding model would be required for which the star-circuit is no longer valid [109]. It was therefore yeasonable to develop another approach, as described here. 6-20 The extension from single-phase to three-phase units turned out to be such easier than was originally thought. Conceptually, each coil of a single-phase unit becomes three coils on core legs I, II, III ina three-phase unit (Fig. 6-9). (a) Three-legged core (b)_ Five-legged core (ec) Shel type design design design Fig. 6.9 - Three-phase transformers In terms of equations, this means that each scalar quantity Z or Y must be replaced by a %3 submatrix of the form ’ (6-28) where 2, {8 the self upedance of the coil on one leg, and Z, 18 the mutual * impedance to the coils on the other two legs on As in any other three-phase network component (e-g-, overhead Line), these self and mutual impedances are related to the positive and zero sequence values, *)prom Pig. 6.9 it is evident that the mutual impedance between legs I and IT is slightly different from the one between legs II and III, etc. Data for this unsyauetry is usually not available, and the unsynmetry is therefore ignored here. To take it into account would require that a three-phase two-winding transformer be modelled as a six-coll transformer (Section 6.4), with 15 measured short-circuit impedances. (6-29) By simply replacing scalars by 3x3 submatrices of the form (6.28), the [R]- and (L]-!-natrix representation of a three-phase transformer 1s found as follows: ne Set up the resistance matrix [R]. If the winding resistances are known, use them in [R]. If they are to be calculated from load losses, use Ea (6-2) for N= 2, or Eq. (6.27) for N= 3. For N> 4, there is no easy way to calculate the resistances. Use positive sequence teat data in these calculations, and assume that the three corresponding coils on legs I, IL, ITE have identical resistances. Find the short-circuit reactances from Eq. (6.25) for positive sequence values. Use the same equation for zero sequence values, provided the zero sequence test between two windings does not involve another winding in delta connection. In the latter case, the data mst first be modified according to Section 6.5.2+ Build the reduced reactance matrix Deere from Eq. (6-20) and (6-22), by first calculating the positive and zero sequence values separately from the positive and zero sequence short-circuit reactances, and by replacing each diagonal and off-diagonal element by a 3x3 submatrix of the form (6.28), ‘The elements of this matrix are calculated with Fa- (6.29). Since the 3x3 submatrices contain only 2 distinct values X, and X,, it is not necessary to work with 3x3 matrices, but only with pairs (X,, X,)- D. Hedman derived a “balanced-matrix algebra” for the multiplication, inversion, etc., of such "pairs" [110], which is used in the support routines BCTRAN and TRELEG. 6-22 4. Invert Laggsveed] to obtain [aresueed) again using Hedman's “palanced-matrix algebra", and expand [ presveed] to the full matrix [B,,] with Eq. (6.24). 5. Since the reactances were in pu+ based on one common S,,,,, the inverse Inductance matrix (1J-! tm actual values 1/H 4s obtained from [B,,] by w §, base multiplying each elenent By, ,, with “z Tyre , where V, and V, are the voltage ratings of coil { and k. For the conversion of p.u. resistances vy, to actual values in@, multiply Ry |, with g+—. a Sbase The procedure of Section 6.5.1 cannot be used directly for the zero sequence calculation of transformers with three or more windings if one or more of them are delta~connected. Assume that a three-winding transformer has wye-connected primary and secondary windings, with their neutrals grounded, and a delta~connected tertiary winding. In this case, the zero-sequence short-circuit test between the primary and secondary windings will not only have the secondary winding shorted but the tertiary winding as well, since a closed delta connection provides a short-circuit path for zero-sequence currents. This special situation can be handled by modifying the short-circuit data for an open delta so that the procedure of Section 6.5.1 can again be used. With the well-known equivalent star cireult of Fig. 6.7, the three test values supplied by the manufacturer are (” subscript dropped to simplify notation), XY, closed A ute x, + ; (6.30a) wee Xup Xyt Xp in peu. values (6.306) ore Xt, (6.30) 6-23 which can be solved for Xys X,s Xp? Xy (6-31a) eerie Ue in psu. values (6.316) Xp = Xue - Xy (6.31¢) After this modification, the short-cireutt reactances Xq +X,» Xy + Xp and X_ + Xp ere used as input data, with winding T no longer being shorted in the test between Hand L. The modification scheme becomes nore complicated if resistances are included. For instance, Eq- (6-30a) becomes G+ 3K) Gy + Rp losed A ; p+ Mh) Oy t Fr =| at By tS ETERS RD | i Pre values (6.32) closed A aE. being the value supplied by the manufacturer, and Ry, Rp, Ry being the winding resistances. This leads to a system of nonlinear equations, which is solved by Newton's method in the support routine BCTRAN. It works for three-winding transformers with wye/wye/delta- and with wye/delta/delta~ connections so far, which should cover most practical cases. 6.6 Exciting Current The exciting current is very auch voltage-dependent above the knee-point™ of the saturation curve \ = f(i). Fig. 6-10 shows a typical curve for a modern high-voltage transformer with grain-oriented steel, with the knee-potnt around 1.1 to 1.2 times rated flux [114]. The value of the Incremental inductance d\/di 1s feirly low in the saturated region, and fairly high in the unsaturated region. The exciting current in the unsaturated region can easily be included in the [L} or ter saturation effects, and extra resistance branches to include excitation -representations. Extra nonlinear branches are needed to include losses. 6-24 6.6.1 For single-phase units and for three-phase units with five-legged core or shell-type design (Fig- 6.9(b) and (c)), the linear exciting current is very small and can often be ignored. If it is ignored, then the [L] matrix representation described in Section 6.3 to 6.5 must be used. A (small) exciting current must alvays be included, houever, {f [L}matrices are used, as explained in Section 6.2. For three-phase units with three-legged core design, the exciting current is fairly high in the zero sequence test (e-g+5 100%), and should therefore not be neglected. The exciting current has an imaginary part, which ts the “magnetizing current” flowing through the magnetizing inductance L,. It also has @ snaller real part (typically 10% of the imaginary part), which accounts for excitation losses. These losses are often ignored. They can be modelled reasonably well, however, with a shunt conductance G, in parallel with the magnetizing inductance L,. The peus magnetizing conductance is G Pexe m pu Satine fee 6-25 and the reciprocal of the p.u. magnetizing reactance is 7 GPE - (6, 5)? (6.34) G pu” Treating with P,,, = excitation loss in excitation test, Ion, = magnitude of exciting current in excitation test, Seating 7 POWEF TAting, and Tearing 7 CUFEERE rating. To assess the relative magnitudes of G, and 1/X,, let us take the values from = 10%, R, = 0.5%, ‘short short, Toxo 7 14)+ Furthermore, assume that the excitation loss V2G, at rated voltage is 25% of the load loss 178, the example of Section 6.2 as typical (x d current (a typical ratio short #¢ Fated current (a typical rat for power transformers). Then G = 0.00125 and I. /T. = 0.01. The im pu exe! Trating reciprocal of the p-u. magnetizing reactance is therefore close to the value of the p.u. exciting current, 1 Texe ety (6.35) fa pu Trating with the error being less than 1% in the numerical example. How to include the Linear exciting current in the model depends on whether an (LJ! - or [L.J-matrix representation ts used, and whether the transformer is a single-phase or a three-phase unit. 6-6.1-1 Single-Phase Transformers In the [L]-matrix representation, the magnetizing inductance 1, will already have been included in the model. Usually, the T-circuit of Fig. 6.5, or the star circuit of Fig. 6.7 with yy connected to star point S, is used in the derivation of [L]. Since L, ,, 18 much larger than Lyyoey pyr tt could be placed across the terminals of the high, low or tertiary side with equal Justification. Alternatively, 2g 4, could be comected to both high and low side, which would convert the T-cireuit of the two-winding transformer into a weireult, or 3L, ,, could be connected to all 3 sides in the case of a 6-26 three-winding transformer. The conversion of L, ., into actual values ts done in the usual way by using the voltage rating for that side to which the inductance is to be connected. For example, connecting the pu. inductance BL, py t0 all 3 sides would mean that the actual values of these 3 inductances are 2 = 3, ‘e am pu Sting en poet Lp * bog FO Tt Pe Se acing In the [L]-4-matrix representation, the “internal” nodes of the T- or star circuit are not available, and the magnetizing inductance mst therefore be connected across one or all “external” terminals, as discussed above. Connecting it across side i is the same as adding 1/L, to the i-th diagonal element of [L]~!. This makes [L]~! nonsingular, and it could therefore be inverted if the user prefers [R]- and [L]-mitrices. This inversion option is available in the support routine BCTRAN, even though this writer prefers to work with [L]-! because [L] is more or less ill-conditioned as discussed in Section 6.2.2. While L, does not create extra branches, but “disappears” instead into the [L]- or [L]-I-matrix, one or more extra resistance branches are needed to model excitation losses with G, ., from Ey. (6-33). Again, G, ,, can either be added to one side, or 7G, pu © both sides of a two-winding transforaer and 5G, p, to all three sides of a three-winding transformer. The conversion to actual values is again straightforward, and R, = 1/G, is then used as input data for the extra resistance branch. 6.6.1.2 Three-Phase Transformers The inclusion of the linear exciting current for three-phase units is 6-27 basically the same as for single-phase units, except that G, and 1/X, from Eq- (6.33) and (6.34) are now calculated twice, from the positive as well as from the zero sequence excitation test data. The reciprocals of the two magnetizing inductances, Boe 7 ML B= 1/t, pos ‘m-pos * ®zero ‘m-zero are converted to a 3x 3 matrix 8 8, 5 BB where B= 4B. + 280.) 3 Prero * pos?» oe, 1 - 3,75 “Flere ~ Boon? * which {s added to the 3x 3 diagonal block in [t]-! of the high, low, or sone other side. Alternatively, }-tiues the pea. 3% 3 matrix could be added to the 3% 3 diagonal blocks of all sides of an N-winding transformer, after conversion to actual values with the proper voltage ratings. After these additions, [t]-! becomes nonsingular and can therefore be inverted for users who prefer [L]-matrices. Support routine TRELEC builds an [L]-matrix directly from both the short-circuit and excitation test data, as briefly described in Section 6.10.3. To include excitation losses, three coupled resistance branches must be added across the terminals of one side. The diagonal and off-diagonal elements of this resistance matrix are L 2 ae oF , freee | “arpos (6.37) Ly Ra ; & m-zero Sm-pos The excitation test for the positive sequence is straightforward, and 6-28 the data is usually readily available. Some precautions are necessary with the zero sequence test data, if it is available, or reasonable assumptions must be made if unavailable. If the transformer hi delta-connected windings, the delta connections should be opened for the zero sequence excitation test. Otherwise, the test really becomes a short-circuit test between the excited winding and the delta-connected winding. On the other hand, if the delta is always closed in operation, any reasonable value can be used for the zero sequence exciting current (e-g-, equal to positive sequence exciting current), because its influence is unlikely to show up with the delta-connected winding providing a short-circuit path for zero sequence currents. If the zero sequence exciting current is not given by the manufacturer, a reasonable value can be found as follows: Imagine that one leg of the transformer (A in Fig. 6.11) is excited, and estimate from physical reasoning how much voltage will be induced in the corresponding coils of the other two legs (B and C in Fig. 6.11). For the three-legged core design of Fig. 6.11, approximately one half of flux h, returns through phases B and C, which means that the induced voltages V, and V, will be close to 0.5 V, (with reversed polarity). If k ts used for this factor 0.5, then 1, exc-zero | 1 +k ; } (6-38) ‘exc-pos = 2 a approx 5 A, approx 5, Fig. 6.11 - Fluxes in three-legged core-type design 6-29 Eq. (6-38) is derived from y= %q Ty (6.39) Vieee cei. (6.39b) with Z,, Z, being the self and mutual magnetizing impedances of the three excited coils. With z, Zooey 7 2, ‘zero — “pos Vy = “KY, omar (6.40) 8 c cents 27 zero * ““pos and Z inversely proportional to I e 5G » Bq. (6.38) ‘pos? “zero ‘exe~pos? Texe-zero follows. Obviously, k cannot be exactly 0.5, because this would lead to an infinite zero sequence exciting current. A reasonable value for I... arg Hi a three-legged core design might be 100%. If I, were 0.5%, k would become 0.49626, which comes close to the theoretical liait of 0.5. Exciting the winding on one leg with 100 kV would then induce voltages of 49.6 kV (with reversed polarity) in the windings of the other two legs. For the five-legged core-type design of Fig. 6.9(b), maybe 2/3 of approximately (1/2)A, would return through legs B and C. In that case, k would be 1/3, oF Teyc-zero/Texe-pos ” 4* The excitation loss in the zero sequence test is higher than in the positive sequence test, because the fluxes X4, hg, Ag in the three cores are now equal, and in the case of a three-legged core-type design must therefore return through air and the tank, with additional eddy-current losses in the tank. Neither the value of the zero sequence exciting current nor the value Of the zero sequence excitation loss are critical if the transformer has delta-connected windings, because excitation tests really become short- circuit tests in such cases. The modification of [1]~! for magnetizing currents and the addition of resistance branches for excitation losses create a model which reproduces the original test data very well. Table 6.1 compares the test data, which was used to create the model with the support routine BCTRAN, with steady-state EMTP solutions in which this model was used to simulate the test conditions (e-g-, voltage sources were connected to one side, and another side was shorted, to simulate a short-circuit test). In this case, the three winding 6-30 resistances were specified as input data, and an [L]-matrix with 10-digit accuracy was used to minimize the problem of ill- conditioning. The excitation data was specified as being measured from the primary side, but 1/L, and shunt conductance G, were placed across the tertiary side, for reasons explained in Section 6.6.2. BCTRAN modifies 1, and R, in this situation, to account for the influence of the short-circuit impedance between the primary and tertiary side. For the zero sequence short-circuit impedance between the primary and secondary side, the modifications of Section 6.5.2 were applied to account for ‘the effect of the delta-connected tertiary winding. 6.6.2 Saturation Effects For the transient analysis of inrush currents, of ferroresonance and of similar phenomena it is clearly necessary to include saturation effects. Only the star circuit representation in the BPA EMTP ("saturable transformer component") accepts the saturation curve directly, while the [L]- and [L]~+-representations require extra nonlinear inductance branches for the simulation of saturation effects. Nonlinear inductances of the form of Fig. 6.10 can often be modelled with sufficient accuracy as two-slope piecewise linear inductances. Fig. 6.12 shows two- and five-slope piecewise linear representations from a practical case [80] for the system shown before in Fig. 6.1. The simulation results (Fig. 6.13) are almost identical, and agree reasonably well with field test results (Fig. 6.14). The slope in the saturated region above the knee is the air-core inductance, which is almost linear and fairly low compared with the slope in the unsaturated region. Typical values for air- core inductances are 2Lgnort (Ugnore= ShOrt-circuit inductance) for two-winding transformers with separate windings [111]. | CIGRE guidelines [200] mention 1 to 1.5 times Lgpore for transformers with Separate windings, and 3 to 4 times Lgnore for autotransformers (= 4 to 5 times Lg,or, seen from the high-voltage side if the nonlinear inductance is'Gonnected across the low-voltage winding). In the unsaturated region, the values can be fairly high on very large transformers (see Fig. 6.10). While it makes little difference to which terminal the unsaturated inductance is connected, it may make a difference for the saturated inductance, because of its low value. 6-31 Table 6-1 - Data for three-phase three-winding transformer in Yyd-connection TYPE OF TEST test pata | SIMULATION RESULTS pos. sequence exciting current (2) 0.428 0.4281 (in phase A) excitation test 0.4280 (in phase B) 0.4280 (in phase C) excitation loss (kW) 135.73 135.731 zero sequence | exciting current (Z) 0.428 | 0.4280 in all excitation test*) phases excitation loss (ki) | 135.73 135.731 short-ctreutt zRS® (%) (300) 8.74 8.740 test impedances, with three-phase | 293° (x) (76) 8.68 8.680 MVA base in parenthesis 2B3° (2) «76 5.31 5.310 235° (xy) (300) | 7.343194"**) | 7.34318 25gT° (*) (300) |26.258183"**) | 26.25806 258°? (2) (300) |18.552826"**) | 18.55284 *) with open delta on gide 3, (values were unavaflable from, test; since, they are unimportant if delta is closed in operation, as explained in text, the positive sequence values were used for zero sequence as well). With closed delta on side 3. These values were calculated from the original test data given as R and X in percent with an accuracy of 2 digits after the decimal point. ey aK) Ideally, the nonlinear inductance should be connected to a point in the equivalent circuit where the integrated voltage is equal to the iron-core flux. To identify that point is not easy, however, and requires construction details not normally available to the system analyst. For cylindrical coil construction, it can be assumed that the flux in the winding closest to the core will mostly go through the core, since there should be very little leakage. This winding is usually the tertiary winding in three-winding transformers, and in such cases it is therefore best to connect the nonlinear 6-32 dvs) 750 Solid line is 5-slope representation; dashed line is simplified 2-slope saturation curve with wh,= 174000 9, wL,= 403.9, X= 775 vs 500 250 air core inductance a ° 20 40 60 icy — Fig. 6-12 - Two-slope and five-slope piecewise linear inductance Fig. 6.13 - Superimposed EMTP simulation results with two~ and five-slope piecewise linear inductances 6-33 SENDING END VOLTAGE (kV) 8 computation field tests 8 RECEIVING END VOLTAGE (kV) computation field tests Fig. 6.14 - Comparison between simulation and field test results inductance across the tertiary terminals. Fig. 6.15 shows the star circuit derived by Schlosser [112] for a transformer with three cylindrical windings (1 closest to core, H farthest from core, L in between), where the integrated voltage in point A is equal to the flux in the iron-core. The reactances of -0.58 @ between A and T is normally not known, but it is so small compared to 7.12 @ between $ and T, that the nonlinear inductance can be connected to T instead of A, with little error. Fig. 6-15 also identifies @ point B at which the integrated voltage is equal to yoke flux. Zikherman [113] suggests to connect another nonlinear inductance to that point B to represent yoke saturation. Since -4.9 @ between H and B is small compared to 22 Q between Hi and S, this second nonlinear inductance could probably be connected to H without too much error. The knee-point and the slope in the saturated region 6-34 Fig. 6.15 - Reactances (in 9) of a three-winding transformer (from [112], which provides the data for 5 cylindrical windings; the two windings farthest from the core are ignored here) of this second nonlinear inductance are higher than those of the first nonlinear inductance (Fig- 6-16). Since it 1s already difficult to obtain saturation curves for the core, this secondary effect of yoke saturation is usually ignored. Dick and Watson [114] came to similar conclusions about the proper placement of the nonlinear inductance when they measured saturation curves on a three-winding transformer. Table 6-2 compares the air-core Inductance (= slope in saturated region) obtained from laboratory tests with values obtained from the star circuit”) if the nonlinear inductance is connected to the tertiary T, or to the star point S. The authors also show a more accurate equivalent circuit which would be useful if yoke saturation or unsymuetries in the three core legs are to be included. If L, is connected to T, then the differences are less than 45%, whereas the differences become very large for the connection to S- Unfortunately, the built-in saturation curve in the BPA star-circuit representation ("saturable transformer component”) is always connected to the star point. This model could become more useful 1£ the code were changed so that L, could be connected to any terminal. this star circuit also had a zero sequence inductance of 1-33 p.u- connected to the high side (see Section 6.6.2+2)+ 6-35, | ay a @u) ee) A, [> " % a 2 Yost i (peu) —> Fig. 6.16 - Nonlinear inductances connected to H (yoke saturation) and L (core saturation) of a two-winding transformer. Reprinted with permission from [113], Copyright 1972, Pergamon Journals Ltd The proper placement of the nonlinear inductance may or may not be important, depending on the circumstances. For example, if the transformer of Table 6.2 with L, in § were energized from the high side, then the amplitude of the inrush current would be correct. If it were energized from the tertiary side, however, then the anplitude of the inrush current would be 56% too low for high levels of saturation”). If details of the transformer construction are not known, then it is not easy to decide where to place 1. In the example of Fig. 6-12-6.14, no construction details were known, and L, was simply placed across the high voltage terminals. In spite of this, simulation results came reasonably close to field test results. 6.6.2.1 Single-Phase Transformers 1f the [1]-!-model of Section 6.3 or 6.4 1s used without the corrections for linear exciting current described in Section 6.6.1, then the nonlinear inductance 1s simply added across the winding closest to the core. If the [u]-wodel of Section 6.2 is used, or if [L]-! nas already been corrected for the Linear exciting current, then a modified nonlinear inductance must be added in which the unsaturated part has been subtracted out (Fig. 6-17). This modified nonlinear inductance has an infinite slope below the knee-point. *)inrush current approximately proportional to 1/L, for flux above ‘air-core knee-point if unsaturated L >> L . 2” ‘atr-core 6-36 Table 6.2 - Comparison between measured and calculated air-core inductances. © 1981 IEEE air-core inductance (p.u.) flux lexctted | measured calculated | error | calculated | error winding | at *) test | withL,inT| (%) | withL, ins} (%) a a 0.207 +465 0.198 0.0 L 0.129 + 4.0 0.120 = 3.2 T 0.076 0.0 0.120 | 458.0 L u 0.129 +16 0.120 7 55 L 0.125 - 48 0.116 -11.0 T 0.076 = 2.6 0.120 +54.0 T a 0.076 0.0 0.120 | 458.0 L 0.076 0.0 0.120 $58.0 T 0.076 0.0 0.173 [4128.0 *) Measured by integrating the voltage at that terminal. The measured short-circuit inductances were Im = 0.0738 p.u., wp = 0.1305 p.U., Ing = 0.0493 p.u., which produces the star- circuit inductances of Ly = 0.0775 p.u., ly = -0.0037 p.u., Ly = 0.0530 p.u. 6.6.2.2 Three-phase transformers For three-phase transformers, the branch matrices [Z] or [¥] reproduce the short-circuit impedances in a "black box" fashion seen from the outside terminals. They contain no information about the internal construction of the transformer. This causes problems when nonlinear inductances are to be added to the [Z] - or [¥] - matrix for the simulation of saturation effects. The proper way for modelling saturation is to look at the magnetic circuit, with its core legs or limbs, yokes, and tank. This magnetic circuit has a dual electric circuit, which can be used to include saturation effects based on the physical details of the construction [121]. Saturation effects are not associated with positive or zero sequence, but with fluxes in the cores, yokes, and tank, though the level of ‘saturation in each part may be influenced by whether excitation is positive or zero sequence. 6-37 To add nonlinear inductances to the "black-box" [Z] - or [¥] - matrix is probably still reasonable for three-legged core designs (Fig. 6.9(a)), where the flux through the winding closest to the core should be more or less identical with the core flux. One could even modify the nonlinear inductance of the two outer cores II and III to take the longer path through half of the bottom and top yokes into account. This is reasonable as long as the three fluxes 4, 4,;, Azz add up to zero. For zero sequence excitation, this is no longer the case. Instead, the fluxes are roughly equal in the three legs, and must therefore return outside the windings through an air gap, structural steel and the tank. Fig. 6.18 shows the measured zero sequence magnetization curve for the transformer described in Table 6.2 [114]. Because of the air gap, this curve is not nearly as nonlinear as the core saturation curve of Fig. 6.10. It is therefore reasonable to approximate it as a linear magnetizing inductance. In [114] it is suggested to connect this zero, sequence magnetizing inductance to the high side. With the [L}"+-model, this is accomplished by setting B,,, = 0 and using Byer, = 1/Lzero in Eq. (6.36), and by adding the 3*x 3 matrix with B, = Ms Brero/3 to the 3 x 3 diagonal block of the high side”). This “puries" the zero sequence magnetizing inductance in [L]~}. If the transformer has delta~connected windings, then any zero sequence excitation is really a zero-sequence short circuit. Zero sequence saturation need therefore not be represented in such cases. A Fig. 6.17 - Subtraction of linear (unsaturated) part in saturation curve (value of A equal in both curves) 2 By setting B,, = 0, [L]~! will remain singular. This causes no problems it the inverse inductance is used. Users who prefer [L]- matrices would have to add another 3 x 3 matrix with B, = 2B,,,/3 and By = ~Byog/3 to one of the sides, with Bo, = 1/Lyog, where Logg is the ifmear (unsaturated) positive "Sequencé* magnetizing inductance. 6-374 (PER PHASED Fig. 6.18 - Zero sequence magnetization curve [114]. © 1981 TEEE 6-38 For five-legged and shell-type designs it is unclear at this time whether saturation representation identical to that for three- legged core transformers discussed above produces acceptable answers or not, or whether the physically-based models of [121] and other references are needed instead. 6.6.3 Hysteresis and Eddy_Current Losses The excitation losses obtained from the excitation test are mostly iron-core losses, because the I*R-losses are comparatively small for the low values of the exciting current. These iron-core losses are sometimes ignored, but they can easily be approximated with the linear shunt conductance G, of Eq. (6.33). A linear shunt conductance G, cannot represent the iron-core losses completely accurately. These losses consist of two parts, Piron-core = Physteresis + Peday current (6.41) namely of hysteresis losses Phyeteregis and of eddy current losses Peaay current: In the excitation tests, these two parts cannot be separated, and only the sum Pi,o,-core iS obtained. Before discussing more accurate representations, it is useful to have some idea about the ratio between the two parts. Ref. [51], which may be somewhat outdated, gives ratios of Physteresis/Peady current = 3 for silicon steel, Physteresis/Peddy current = 2/3 for grain-oriented steel, while a more recent reference [125] quotes a typical ratio of 1/3. On modern transformers, hysteresis losses are therefore much less important than they used to be before the introduction of grain- oriented steel. It is generally agreed that eddy current losses are Proportional to 4? and to f? [51], at least in the low-frequency range, which seems to change to f1:5 in the high-frequency range because of skin effect in the laminations. Frequency-dependent eddy current representations were discussed in [115], where R, is replaced by a number of parallel R-L branches. It is doubtful whether this sophistication is needed, however, because the reduction caused by a proportionality change from f* to f!-> at high frequencies is probably offset by other types of loss increases (e.g., by increases in coil 6-39 resistance due to skin effect, etc-)- At any rate, laboratory tests would first have to be done to verify the correctness of the frequency dependence proposed in [115]. In such tests it may be difficult to separate eddy current and hysteresis losses. If we accept a proportionality with \? and 2, then a constant resistance R, does model these losses very well, because Poady current ~ YRMs/®y 24 Vays 707 Mays for sinusoidal excitation. Hysteresis losses are a nonlinear function of flux and frequency, a > Paysteresis ~*A) * Cf)" ue In [51], a 1s said to be close to 3 for grain-ortented steel, and b= 1, In [116], a= 2.7 and b= 1.5. If a= b = 2 were used, then the sum of hysteresis and eddy current losses could be modelled by the constant resistance R, or conductance G, of Eq- (6-33). This ie a reasonable first approximation [125], especially if one considers that hysteresis losses are only 25% of the total iron-core losses in transformers with grain-oriented steel. Pig. 6.19(a) shows the nonlinear inductance of a current transformer, which was used by C. Taylor to duplicate field test results in a case where the secondary current was distorted by saturation effects [117]. Fig. 6-19(b) shows X as a function of the exciting current in the transient simulation, 1f tron-core losses are modelled with a constant resistance R, = 800. It can be seen that R, not only creates the typical shape of @ nornal magnetization curve (with lower d\/di coming out of the origin, compared to = £(4) in Fig. 6.19(a)), but also creates minor loops with reasonable shapes. If the flux-current loop’) for sinusoidal excitation is available, then R, can also be calculated from oe 6.43) as an alternative to Ba. (6.33), with Bt being half of the horizontal wideh of the loop at A= 0 (Fig. 6-20), and v= UA... Eqs (6.43) is derived from *)the author is reluctant to call it “hysteresis loop” because the losses associated with this loop are the sum of hysteresis and eddy current losses, with the latter actually being the larger part in transformers with grain-oriented steel. 6-40 008 008 ! A (e/earm) 2 (wa/eemm) 006 | 006 | | Foxcseing | 004 +004 R= 160 2 fe ey 002 002 6 7 Es eo w % % 2 edo toms) exciting tm) (a) Nonlinear magnetizing inductance (b) Loops created by constant of current transforner for hysteresis and eddy current losses Fig. 6.19 - Saturation in current transformer [117]. Reprinted by permission of C.W. Taylor realizing that at A = 0 all the current must flow through the parallel resistance R, and that the voltage reaches its peak value wA,,, at 4 = 0 because of the 90° phase shift between voltage and flux. If more values of Ai are used at various points along the A-axis, together with the corresponding values for v = dk/dt, then a resistance R, can be constructed which becomes nonlinear. This parallel combination of nonlinear resistance and nonlinear inductance has been proposed by L.0. Chua and K.A. Stromsmoe [118] to model flux-current loops caused by hysteresis and eddy current effects. They give convincing arguments why this representation is reasonable. In particular, they did make comparisons between simulations and laboratory tests, not only for a small audio output transformer with laminated silicon steel, but for a supermalloy core inductor well. Fig. 6.21 shows the nonlinear inductances and resistances for this audio output transformer [118]. Fig. 6-22 compares the laboratory test results with simulation results [118] (first row laboratory results, second row simulation 6-41 results). Fig. 6.22(a) is a family of flux-current loops for 60 Hz sinusoidal flux linkage of various amplitudes. Fig. 6-22(b) shows two loops, one with a sinusotdal flux Linkage and the second with a sinusoidal current. Fig. 6-22(c) 1s a family of loops obtained at 60 Hz for various amplitudes of sinusoidal current. Fig. 6.22(d) shows a family of loops for sinusotdal flux Fig. 6-20 - Flux-current loop Linkages at 60, 120, and 180 Hz. In all cases, the agreement between measurements and simulation results is excellent. The minor loops in Fig. 6.22(e) were obtained with a 60 Hz sinusoidal current superimposed on a de bias current. Again, there appears to be excellent agreement. The aajor dravback of this core-loss representation with a linear or nonlinear resistance ts its inability to produce the correct residual flux when the transformer is switched off. This was one of the motivations for the development of more sophisticated hysteresis models, but even these models do not seem to produce the residual flux very accurately. This writer believes that there are no models available at this time which can predict residual fluxes reliably, and that reasonable assumptions should therefore be 6-42 i (ma) — Fig. 6.21 ~ Model for exciting current with parallel, nonlinear resistances and inductances [118]. © 1970 IEEE made. There is no difficulty with the linear or nonlinear R,-representation in starting a transient simulation with a residual flux if its value is Provided as input data, as explained in Section 6.6.4. The more sophisticated models mentioned above use pre-defined trajectories or “templates” in the \, plane to decide in which direction the curve will move if the flux either increases or decreases [114,119]. The technique of [119] has been implemented in the BPA-EMIP (“pseudononlinear hysteretic reactor”) but a careful comparison with the simpler R -representations (either linear or nonlinear) has not yet been done. More research may be needed before reliable hysteresis models become available. Such models may be based on the duality between magnetic and electric circuits, which would then require the dimensions of the iron-core as input data [121], or they may be based on the physics of magnetic materials [120]. 6-43 (a) (ai (e) Fig. 6.22 - Comparison between measured and simulated flux-current loops [118]. © 1970 TEEE 6.6.4 Residual_Flux Residual flux is the flux which remains in the iron core after the transformer 1s switched off”). Ie has a major influence on the magnitude of inrush currents. Starting an EMTP simulation from a known residual flux is relatively easy, with simple as well as with sophisticated hysteresis aodels. To find the residual flux from a simulation is more complicated, and the results still seem to be unreliable at this time, even with sophisticated *) there seems to be some confusion in terminology between “residual” and “remanent” flux. It appears that remanent flux is the flux value at i=0 in the hysteresis curve under the assumption of sinusoidal excitation. 6-44 hysteresis models. Until this situation improves, it might be best to use a typical value for the residual flux as part of the input data. Unfortunately, not much data is available on residual flux. A recent survey by CIGRE [122] has not added mich to it either, except for the quotation of 2 maximum values of 0.75 and 0.90 p.u. This survey does contain a reasonable amount of information about values of air-core inductances and saturation curves, however. ‘The UBC version of the EMTP starts the simulation from a nonzero residual flux with the following approach, in connection with piecewise Figure 6.23 - Starting from residual flux linear inductances"") (see also Section 12.1.3): At tO, the starting point Alles at Apegquar aM¢ 190, and the simulation moves along a slope of 1, (unsaturated value), as shown in Fig. 6.23. The alone is changed to L, (saturated value) in point B as soon as 4 > Ayggr AE the sane time, a value Aguttch #8 calculated which will bring the characteristic back through the **)in the BPA version, this branch type has been generalized from 2 to n slopes ("pseudononiinear inductor”), but it appears that it no longer accepts residual flux as input data. 6-45, origin when the slope is changed back to L, as soon as A < switch? Thereafter, the normal A/i-curve will be followed. More details, in particular the problem of overshoot (A slightly larger than A... when going into saturation, and A slightly lower than A,,4,,), when coming out of saturation), are discussed in Section 12.1.3.3. For typical saturation curves, such as the one shown in Fig. 6.10, the linear slope is almost infinite; in that case, the first move into saturation practically lies on the given A/i-curve, rather than somewhat higher as in Fig. 6.23. The simple hysteresis model of a nonlinear L in parallel with a resistance R, cannot be used to predict the residual flux after the transformer is switched off. The energy stored in L, will simply be dissipated in R, in this model, with an exponential decay in current and flux to zero values. The flux value at the instant of switching could possibly be close to the residual flux, but this has never been checked. Also, this value would only be meaningful if the transformer is switched off by itself, without lines or other equipment connected to it. 6-7 Autotransformers Tf an autotransformer is treated the same way as a regular transformer, that is, if the details of the internal connections are ignored, the models discussed here will probably produce reasonably accurate results, except at very low frequencies. At dc, the voltage ratio between the low and high side of a full-winding transformer will be zero, whereas the voltage ratio of the autotransformer of Pig. 6.24 becomes Ry;/R, (de voltage divider effect). For a more accurate representation, series winding I and common winding II should be used as building blocks, in place of high side H and low side L. This requires a re-definition of the short-circuit data in terms of windings I and II. Since most autotransformers have a tertiary winding, this winding T shall be included in the re-definition. First, the voltage ratings are 6-46 1 L Gao, (6.44) Vu 7 Yr The test between H and L provides the required data for the test between I and II directly, since IT {s shorted and since the voltage applied to H is actually applied to I (b and c are at the sane potential through the short-circuit connection). Only the voltage ratings are different, and the conversion from H to I is simply in peu. values (6.45) No modifications are needed for the test between II and III, Zr” Aur in peu. values (6.46) H b a |u xr . Fig. 6.24 - Autotransformer with tertiary winding For the test between H and T, the modification can best be explained in terms of the equivalent star-circuit of Fig. 6.7, with the impedances being in this case. With IIT short- try 7 Vp) will flow through 257+ current will also flow through I and II as 1 p.u. based on V,, or converted L Zys Zyps Zyzz> based on Vy, Vyye Very circuited, 1 p-u- current (based on V, This to bases Vy, 17 Wq 7 Vp)/Vy and Ty, = Vp/Vy+ With these currents, Yup the p.u. voltages become 6-47 172 yt a + in peu. values (6.47) L “a 4 et a * in peu. values (6-48) Converting V; and V,, to physical units by multiplying Eq. (6.47) with (Wq 7 Vp) and Eq. (6.48) with V,, adding them, and converting the sum back to a peu. value based on V,, produces the measured p.u. value Va ~ V2 2 y, Ly +2 Ge + Zz, im peu. values (6.49) Eqs- (6.45), (6.46) and (6.49) can be solved for 21, Zyys Zyzy since 2111 = 2] + 2, and 211,111 "2+ 2p Va Vy Saran? 2a, ES in peu. values (6-50) q - Y The autotransformer of Fig. 6.24 can therefore be treated as a transformer with 3 windings 1, II, ILI by simply re-defining the short-circuit impedances with Eqs. (6-45), (6-46) and (6.50). This must be done for the positive sequence tests as well as for the zero sequence tests. If the transformer has a closed delta, then the zero sequence data must be further modified as explained in Section 6.5.2, after the re-definition of the short-circuit data. 6.8 Ideal Transformer An ideal transformer was not added to the BPA EMTP until 1982. The ideal transformer has no impedances and simply changes voltages and currents from side 1 to side 2 (Fig. 6.25) as follows: u 4 L rasta (6-51) y a? 6-48 1 x i} ‘] “1 2 n 2 Fig. 6.25 - Ideal transformer It is handled in the system of nodal equations (1.8a) or (1.20) by treating current 4, as a variable, and by adding the equation ay, 7 av, = (vy - vy) = 0 (6.52) The matrix of the augmented systen of equations, with an extra column for variable i,, and an extra row for Eq. (6-52), then has the form of Fig. 6.26. column for i, k n normal {G]-matrix a L row for Bq. (6.52) Fig. 6.26 - Augnented [¢]-natrix ‘The ideal transformer can also be simulated with 8 resistance branches and one extra node “extra”, as shown in Fig. 6.27, because these branches 6-49 augment the matrix in the same way as shown in Fig. 6.26. In both approaches it 1e important that node “extra” (or Eq. (6-52)) is eliminated after nodes k, m, J, 2, to assure that the diagonal element becomes nonzero during the elimination process. Fig. 6.27 - Resistance modelling of ideal transformer If the transformer is unloaded (i ), the elimination process will fail with a zero diagonal element. The UBC version would stop in that case with an appropriate error message, while the BPA version will first print a warning, and then continue after automatic connection of a very large resistance to the node where the zero diagonal element has been encountered. This problem 1s related to the treatment of floating subnetworks (see next Section 6.9). 6.9 Floating Delta Connections Most transmission autotransformers have delta-connected tertiary windings for the suppression of third harmonics. Frequently, nothing is connected to such tertiary windings. In that case, and in similar cases, the delta windings have floating potential with respect to ground (Fig. 6-28): only the voltages across the windings a-b, bre, cna are defined, but not the 6-50 voltages in a, b, or c with respect to ground. Since the EMTP solves for node voltages with respect to ground, the Gauss elimination will fail with a zero diagonal element. © > PPPPPPPIPOD Fig. 6-28 - Floating delta connection To prevent the solution algorithm from failing, one can either ground one of the nodes (e-g-, node a), or connect stray capacitances or large shunt resistances to one or all 3 nodes. Connecting identical branches to each of the 3 nodes has the “cosmetic” advantage that the voltages in a, b, ¢ will be symmetrical, rather than one of them being zero. The BPA version connects a large shunt resiatance automatically, with an appropriate warning, whenever a zero or near-zero diagonal element is encountered. For example, if the zero diagonal is encountered at node c, then a large resistance will be connected from c to ground which will make v, = 0. 6.10 Description of Support Routines and Saturable Transformer Component Except for the “Saturable Transformer Component” in the BPA EMTP, which is an input option specifically for transformers, all other transformer representations discussed here use the general branch input option for n-cireutts (with 0-0), and possibly additional linear or nonlinear, uncoupled resistance and inductance branches for the representation of the exciting current. There are three support routines XFORMER, TRELEG and BCTRAN, which convert the transformer data into impedance or admittance matrices, as well as a support routine CONVERT for the conversion of saturation curves Vous 7 nto A = £(i). These euppoi as wel ig 7 FClgys) into A = £4). The pport routines, as well ea the 6-51 built-in saturable transformer component, are briefly described here. 6.10.1 ‘This support routine for single-phase transformers is somewhat obsolete, and has been superceded by support routine BCTRAN. For two-winding transformers, it uses essentially the approach of Section 6.3.1 to form an admittance matrix oe ou Zu Md "fla a |e pee: im without first separating R and L as in Eq. (6.7). One half of from ™ pu Eq. (6.35) 19 then added to Yy, 4, and Ypp ,,» which makes the matrix nonsingular. After its inversion, and conversion from p.u. to actual values, the 2% 2 branch impedance matrix is obtained. By not separating R and L, this impedance matrix has nonzero off-diagonal resistances, which would produce wrong results at extremely low frequencies when the magnitude of R becomes comparable with the magnitude of wL (in one particular example, R= wl at £ = 0,002 Hz). At de, an off-diagonal resistance would imply a nonzero induced voltage in the secondary winding, which should really be zero in a full-winding transformer. For three-winding transformers, the approach of Section 6.3.2 is used. First, the impedances of the equivalent star circuit are found with Eq. (6-10), which is then converted to the delta circuit with Eq. (6-11) to obtain the 3% 3 adaittance matrix [¥,,] of Ba. (6-12). Again, there ts no separation between R and L, and complex impedances 2 are used in place of X from Eq. (6-35) is then added to 1m all these equations. One third of 3 ™ po Yar pur Yaz pu 24 ¥93 pyr followed by matrix inversion and conversion to actual values. Again, nonzero off-diagonal resistances will appear in the branch impedance matrix, as already discussed for the two-winding 6-52 transformer. Except for errors at extremely low frequencies, which is caused by not separating R and L, the model produced by XFORMER is useful if the precautions for ill-conditioned matrices discussed in Section 6.2.2 are observed. 6.10.2 This support routine works for any number of windings, and for single-phase as well as for three-phase units. Tt uses the approach of Section 6.4 and 6.5 to produce the [8] and [t]~!-—matrices of coupled branches. BCTRAN has an option for inductance matrices [1] as well, in cases Where the exciting current is nonzero. Because of the {11-conditioning problen (Section 6.2.2), the author prefers to work with [L]"! instead of [1], however. Impedance matrices produced by BCTRAN and XFORMER differ mainly in the existence of off-diagonal resistance values in the latter case, which should make the model from BCTRAN more accurate than that from XFORMER at very low frequencies. 6.10.3 Support Routine TRELEG This support routine was developed by V. Brandwaja at Ontario Hydro, concurrently with the development of BCTRAN at UBC. Tt builds the impedance matrix (6.14) of N-winding single-phase or three-phase transformers directly from short-circuit and excitation test data, without going through the reduced impedance matrix described in Section 6.4. The exciting current must alvays be nonzero, and for very small values of exciting current, the matrices are subject to the {1l-conditioning problem described in Section 6.2.2. Recall that Eq. (6-14) is valid for three-phase transformers as well, if each element is replaced by a 3% 3 submatrix as discussed in Section 6.5. With this in mind, the imaginary parts of the diagonal elenent pairs (X, 415 Xi of the excited winding "i" are first calculated from the current of 6-53 the positive and zero sequence excitation tests. If excitation losses are ignored, then X,, in per unit is simply the reciprocal of the per-unit at exciting current. With positive and zero sequence values thus known, the pair of self and mutual reactances is found from Eq. (6.29). For the other windings, {t is reasonable to assume that the p.u. reactances are practically the same as for winding “i", since these open-circuit reactances are much larger than the short-circuit impedances. This will produce the imaginary parts of the other diagonal elements”). The real part of each diagonal element is the resistance of the particular winding. With the diagonal element pairs known, the off-diagonal element pairs (Zea Zyagy) Ate calculated from Eq. (6.5), except that real values X are replaced by complex values Z, Bae (yy = HEY Ty (6-53) ‘These impedances are first calculated for positive and zero sequence, and then converted to self and mutual impedances with Eq. (6.29). As pointed out in Section 6.2.2, the elements of [Z] must be calculated with high accuracy; otherwise, the short-circuit impedances get lost in the open-circuit impedances. The lower the exciting current is, the more equal the p.u. impedances Z,,, Z,, and Z,, become among themselves in Eq. (6.5)+ Experience has shown that the positive sequence exciting current should not be much smaller than 1% for a single~precision solution on a UNIVAC computer (word length of 36 bits) to avoid numerical problems. On computers with higher precision, the value could obviously be lower. On large, modern transformers, exciting currents of less than 1% are common, but this value can usually be increased for the analysis without influencing the results. Since these 111-conditioning problems do not exist with [1]-!, support routine BCTRAN should make TRELEG unnecessary, after careful testing of both routines has been carried out. “If it is known that the magnetizing impedance should be connected across a particular terminal, then the diagonal elements are modified to account for the differences caused by the short-circuit impedances between the terminals. 6-54 Often, saturation curves supplied by manufacturers give RMS voltages as * ‘a function of RMS currents. The support routine CONVERT” Vays! TpygnCurves into flux/current-curves \ = £(4) with the following changes simplifying assumptions: 1. Hysteresis and eddy current losses in the iron-core are ignored, 2. resistance in the winding is ignored, and 3. the A/i-curve is to be generated point by point at such distances that linear interpolation is acceptable in between points. For the conversion it is necessary to assume that the flux varies sinusoidally at fundamental frequency as a function of time, because it is most Iikely that the Vayc/Tgyg-curve has been measured with a sinusoidal terminal voltage. With assumption (2), v= d\/dt. Therefore, the voltage will also be sinusoidal and the conversion of Vays Values to flux values becomes a simple re~scaling: aus!” 2 Ae (6.54) ‘The re-scaling of currents is more complicated, except for point i, at the end of the linear region A-B (Fig. 6.29): ty = Tigo’? (6-55) are found recursively: Assume that {, 1s the next value to be found. Assume further that the sinusoidal flux Just reaches The following points ig, tps the value h, at its maximum, Awdg sin ut « (6.56) Within each segment of the curve already defined by its end points, in this case A-B and B-C and C-D, i is known as a function of \ (namely piecewise linear), and with Eq. (6-56) is then also known as a function of time. Only the last segment is undefined inasmuch as i, is still unknown. Therefore, *)convERT was developed with the assistance of C.F. Cunha, CEMIG, Belo Horizonte, Brazil. linear intexpolation between points only discrete points used ae S Loyg — 2 os RMS Fig. 6.29 - Recursive conversion of @ Vays/Tpygcurve into a h/i-curve inf(t,i,) tn the last segment. If the integral needed for RMS-values, x ae Peis, Pawey (6.57) is evaluated segment by segment, the result will contain 1, as an unknown variable. With the trapezoidal rule of integration (reasonable step size = 1°), F has the form 5 2 Fea+ di, + ct (6.58) with a, by © know. Since F must be equal to Ipc p by definition, Eq. (6.58) can be solved for the unknown value ig. This process is repeated recursively until the last point iy has been found. If the A/i-curve thus generated is used to re-compute @ Vayo/Tpyscurve, it will match the original Vays/Tpygcutve, except for possible round-off errors. As an example, upport routine CONVERT would convert the table of per-unit RMS exciting currents as a function of per-unit RHS-voltages, with base power = 50 MVA and base voltage 635.1 kV, into the following 6-56 flux/current relationship: As) Aca) 0 0 2144.22 0.6235 2382.46 2.7238 2620.71 7.2487 This A/i-curve is then converted back into 4 Vays/Tpyg"curve as an accuracy check. Tn this case, the Vays and Tpyg Values were identical with the RMS S original input data. Very often, the Viyc/I Pacman ‘ans! ‘RMS: not for high values of saturation. In such cases, it is best to do the -curve is only given around the knee-point, and conversion first for the given points, and then to extrapolate on the A/t-curve with the air-core inductance. 6.10.5 Saturable Transformer Component This built-in model was originally developed for single-phase N-winding transformers. It uses the star-circuit representation of Fig. 6.30. The primary branch with R,, 1, is handled as an uncoupled R-L branch between nodes BUSL,, and star point S, whereas each of the other windings 2,...N is treated as a two-winding transformer (first branch from $ to BUS2,, second branch from BUSI, to BUS2,, with k=2,...N). The equations for each of these two-winding transformers are derived from the cascade connection of an ideal transformer with an R-L-branch (Fig. 6.31). This leads to (ye x 9 tocar] a | om im u ‘oear Sy yae MS |[se3 =n 7 = | lee a uy (6.59) 6-57 BUS1, winding 1 Bus2. ideal Fig. 6.31 - Cascade connection of ideal transformer and R-L-branch which is the alternate equation (6.1) with an inverse inductance matrix {L]-*. In the particular case of Eq. (6.59), the product [L]-1[R] is symmetric, which is not true in the general case. [L]-1 from Bq. (6.59) with [R]= [s z, | produces an unsymmetric matrix, o Me ty-1iy = 2 a, & |. 1 os Actually, miltiplying 6-58 which can be brought into the symmetric form of Eq. (6.59) by substituting 7 i, into the first row, This can only be done if there is no "1 magnetizing inductance, as in Fig. 6.31. ‘star The input data consists of the R, L-values of each star branch, and the turns ratios, as well as information for the magnetizing branch. For three-winding transformers, the impedances of the star branches are usually available in utility companies from the data files kept for short-circuit studies. If these values are in p.u., they must be converted to actual values by using the proper voltage rating V, for each of the star branches kel,...N, If the short-circuit impedances are known, then the star branch impedances can be calculated from Eq. (6.10). The saturable transformer component has some limitations, which users should be aware of: 1, It cannot be used for more than three windings, because the star circuit is not valid for N>3. This is more an academic than a practical limitation, because transformers with more than three windings are seldom encountered. 2, The linear or nonlinear magnetizing inductance, with R, in parallel, is connected to the star point, which is not always the best connecting point, as explained in Section 6.6. 3. Numerical instability has occasionally been observed for the three-winding case. It is not believed to be a programming error. The source of the instability has never been clearly identified, though it is felt that it is caused by the accumulation of round-off errors. V. Brandwajn tan a case in 1985 in which the instability disappeared when the ordering of the windings was changed (e.g., first winding changed to low side from high side). 4, While the saturable transformer component has been extended from single-phase to three-phase units through the addition of a zero-sequence reluctance parameter, its usefulness for three-phase units is limited Three-phase units are better modelled with inductance or inverse inductance matrices obtained from support routines BCTRAN or TRELEG. 6-59 6.11 Frequency-Dependent Transformer Models At this time, no frequency-dependent effects have yet been included in the transformer model. There are basically three such effects: a. Frequency-dependent damping in the short-circuit impedances, b. frequency dependence in the exciting current, and c. influence of stray capacitances at frequencies above 1 to 10 kHz. CIGRE Working Groups [9,18] have collected some information on the frequency-dependent L/R-ratios of short-circuit impedances (Fig. 2.17). As explained in Section 2.2.3, this frequency dependence can easily be modelled with parallel resistances, which matches the experimental curves reasonably well (Fig. 2.19). When dealing with matrices [L] or [L]-1, resistance or conductance matrices [R,] or [G,] could be added automatically by the Program, with the user simply specifying the factor k in 2 OR} = KEL), or (6) = ¢ 1) (6.60) Frequency-dependent effects in the exciting current were modelled with parallel R-L branches in [115], as discussed in Section 6.6.3. Whether the linear frequency dependence in these parallel R-L branches can be separated easily from the nonlinear saturation effects would have to be verified in laboratory experiments. For transient studies which involve frequencies above a few kiz, capacitances must be added to the R-L-models. As suggested in [123], capacitances should be included 4, between the winding closest to the core, and the core, b. between any two windings, and €. across each winding from one end to the other. In reality, inductances and capacitances are distributed, but reasonably accurate results, as seen from terminals, can be obtained by lumping one half of the capacitance at each end of the winding for effects (a) and (b), and by lumping the total capacitance in parallel with the winding for effect (c), as 6-60 shown in Fig. 6.32. Each of these capacitances can be calculated from the geometry of the transformer design. Obviously, the internal voltage distribution across a winding, which is of such great concern to the transformer designer, cannot be obtained with the simple model of Fig. 6.32. Fig. 6.33 compares measured impedances of a transformer (500 HVA, 765/345/17.25 kV) and calculated impedances with a model where the capacitances were added according to Fig. 6.32. The agreement is quite good. Similar suggestions for the addition of capacitances have been made by others (e.g., (124). nie oe core oF Fig. 6.32 - Addition of capacitances to R-L-model (subscripts refer to the three effects mentioned in text) 6-61 (a) X1 excited, H1,¥1,¥2 grounded —(b) X1 excited, H1,¥1,¥2 open-circuited Fig. 6.33 - Frequency response of single-phase autotransformer with tertiary winding (marking of terminals according to North American standards: Hl = high voltage terminal, Xl = low voltage terminal, Yl, ¥2 = terminals at both ends of tertiary winding) (123). ©1981 IEEE. roe 7. SIMPLE VOLTAGE AND CURRENT SOURCES Most of the simple sources are either voltage or current sources defined as a time-dependent function f(t), v(t) = £(t), or A(t) = £(t) + (ely Frequently used functions f(t) are built into the EMTP. There is also a current-controlled de voltage source for simplified HVDC simulations, which ie more complicated than Eq. (7-1). In addition to the built-in functions, the BPA version of the EMIP allows the user to define functions through user-supplied FORTRAN subroutines, and to declare TACS output variables as ‘The UBC version of the EMTP does not voltage or current source function: have these two options, but allows the user to read f(t) step by step in Amcrements at At. This option has rarely been used however. Note that £(t) = 0 for a current source implies that the source is disconnected from the network (i=0), whereas for a voltage source it implies that the source is short-circuited (v=0). 7.1 Connection of Sources to Node If a voltage or current source is specified at a node, it is assumed to be connected between that node and local ground, as shown in Fig. 7-1. A voltage source of v(t) = +1-0 V means that the potential at that node is 41.0 V with respect to local ground, whereas a current source of +1.0 A implies that 1.0 A flows from the local ground into that node. vit) _ ine £(e) ig = -£(e) (a) Voltage source between (b) Current source from (c) Current source node and local ground local ground into node between two nodes Fig. 7-1 - Source connections 7-2 7.2 Current Sources Between Two Node: Current sources between two nodes, e-g-, a current leaving node B and entering into node A as shown in Fig. 7-1(c), must be specified as two current sources, namely as Ag(t) = £(t), and 14(t) = -£(t) (7.2) 7.3 Voltage Sources Between Two Nodes Until recently, voltage sources could not be connected between two nodes. With the addition of ideal transformers to the BPA EMTP in 1982 (Section 6.8), voltage sources between two nodes are easy to set up now. In Fig. 6.25, simply ground node 2, connect the voltage source from node j to ground, and use a transformer ratio of 1:1. This will introduce a voltage source between nodes k and m. A special input option has been provided for using the ideal transformer for this particular purpose. The UBC EMTP and older versions of the BPA EMTP do not accept voltage sources between nodes. One could use the equivalent circuit of Fig. 6.27 for the {deal transformer, however, which turns into the circuit of Fig. 7.2. This representation works in the transient solution part of the UBC ENTP, provided the branches of Fig. 7-2 are read in last. In that case, the node “extra” will be forced to the bottom of the equations as shown in Fig. 6.26. The steady-state subroutine in both versions, as well as the transient solution in the BPA version, use optimal re-ordering of nodes, which may not force the row for node “extra” far enough down to assure nonzero diagonal elements during the Gauss elimination. Using Fig. 7-2 may therefore not always work, unless minor modifications are made to the re-ordering subroutine. In all versions, a voltage source in series with a (nonzero) impedance can always be converted into a current source in parallel with that impedance. The current source between the two nodes is then handled as shown tn Eq. (7-2). The conversion from a Thevenin equivalent circuit (v in series with Z) to a Norton equivalent circuit (1 in parallel with Z) is especially simple {£ the impedance is a pure resistance R, as shown in Fig. 7.3. -19 vit) extra ¢ 4— itt) = $9) 1a -12 Pig. 7.2 Equivalent circuit for voltage source v(t) between nodes k and m 7 KY’ ke’ v(t) convert to: i(t) ue I R my my Fig. 7.3 - Conversion of v(t) in series with R into i(t) = v(t)/R in parallel with R Converting a voltage source in series with an inductance L into a current source with parallel L de slightly more complicated. L is again connected between nodes k and m, in the same way as Rin Fig. 7.3. The definition of the current source depends on the initial conditions, however. For example, if v(t) = Van, cos(wt + o) (7.3) and Lf the case starts fron zero intial conditions, then \ 1(e) = BE [atnuet ¢) ~ ete] (sta) If the case starts from linear ac steady-state conditions, with that voltage source being included in the steady-state solution, then Yaa: A(t) = X coa(utt @ - 90°) « (7.4) T-4 7.4 More Than One Source on yme_Node If more than one voltage source is connected to the sane node, then the EMTP simply adds thetr functions f,(t), £,(t) to form one voltage source. This implies a series connection of the voltage sources between the node and local ground, as shown in Fig. 7.4(a)- t t ' 1 rm (a) Series connection of voltage (b) Parallel connection of current sources sources Fig. 7.4 - Multiple voltage or current sources on same node If more than one current source is connected to the same node, then the EMTP again adds their functions £,(t),++-f,(t) to form one current source. This implies a parallel connection of the current sources, as shown in Fig. 7.4(b)« Source functions can be set to zero by using parameters Toy,yp and Tyyop: The EMTP sets £(t) = 0 for t < Tyragp and for t > Tgrop- By using more than one source function at the same node with these parameters, more complicated functions can be built up from the simple functions, as explained in the UBC User's Manual and in the BPA Rule Book. Té voltage and current sources are specified at the same node, then only the voltage sources are used by the EMTP, and the current sources are 1-5 ignored. Current sources would have no influence on the network in such a case, because they would be directly short-circuited through the voltage sources. 7.5 Built-in Simple Source Functions Commonly encountered source functions are built into the EMTP. They are: (a) Step function (type 11). In cases which start from zero initial conditions, the step function is approximate in the sense that the EMTP will see a finite rise time from f(0) = 0 to f(St) = F,,., a8 shown in Fig. 7.5. 7 ' £(t) at to t= (a) Starting from zero initial (b) Starting from initial value Fag conditions Fig. 7.5 - Step function (b) Ramp function (type 12) with f(t) as shown in Fig. 7.6. The value of the function rises linearly from Trapp © Typagy + Ty tO @ value of Fay, and then remains constant until it is zeroed at t > Type TstaRt Tsrop t Fig. 7.6 - Ramp function 7-6 A modified ramp function (type 13) has the same rise to Fa, at Tape + Ty as in Fig. 7.6, but decays or rises with a linear slope thereafter. By setting Tyran = 0 and Ty = 0, this becomes a step function with a superimposed linear decay or rise. (c) Sinusoidal function (type 14) with £(t) = Fa, cos(ut +4) Mf 7, ° (7.58) ‘START or £(t) = Fagg c08(W(t-Topgag) +9) Lf Tyrapp > 0 (7.50) with £(t) = 0 for t < Topappe This 1s probably one of the most used source functions. Note that the peak value F,, must be specified, rather than the RMS value. To start 4 case from linear ac steady-state conditions, or to obtain a sequence of steady-state solutions at a number of frequencies, use Tgyanq < 0 to indicate to the EMIP that this sinusoidal source should be used for the steady-state solution. The value of 1, START negative, and the complex peak phasor used for that source is then is immaterial as long as ite value is ie VorI* Fae (7.6) (a) Impulse function (type 15) of the form F(t) = K(e TE - e M2") (1) This function has been provided for the representation of lightning or switching impulses, as used in standard impulse tests on transformers and other equipment. A typical lightning impulse voltage is shown in Fig. 7.7 [126], and a typical switching impulse voltage is shown in Fig. 7.8 [126]. There is no simple relationshtp between the time constants 1/a, and 1/a) in Bq- (7.7) and the virtual front time 1 (or time to crest T.,) and the virtual time to half-value T,. Table 7.1 shows the values for frequently used waveshapes, as well as values for k which produce a maximum value of fqax 7 10 in Eq. (7.7). The time at which the maximum occurs is found by 1-1 setting the derivative df/dt = 0 from Eq. (7.7) and solving for t__ Inserting t,,, into Eq. (7-7) then produces £,,.. Note that 1/a, and 1/a, in Table 7.1 are in us, whereas the EMTP input is usually ins v 19} Tail Fig. 7.7 - General shape of lightning impulse voltage (IEC definitions: 1, = virtual front time, typically 1.2 us + 3025 T; = virtual time to half-value, typically 50 us + 20%). Reprinted with permission from [126], © 1984, Pergamon Books Ltd v 10 os} - os. Fig. 7.8 - General shape of switching impulse voltage (IEC definitions: T,, 7 time to crest, typically 250 us + 20%; T, = virtual time to'half-value, typically 2500 us + 60%; Ty = time above 90%). Reprinted with permission from [126], © 1984, Pergamon Books Led In impulse testing, the capacitance of the test object is usually mich smaller than the capacitance of the impulse generator. It is then permissible to regard the impulse generator as a voltage source with the function of Eq. (7.7). In cases where the impulse generator is discharged into Lines, or into other test objects with impedances which can influence 7-8 Table 7.1 - Relationship between T,, T, and 4, a2. Reprinted with permission from [126], © 1984, Pergamon Books Ltd T/% Gs) | T/T (us) 5 (ws) 5 (hs) | to produce max 1.2/5 - 3.48 0.80 2.018 1.2/50 - 68.2 0.405, 1.037 1.2/200 i 284 0.381 1.010 250/2500 - 2877 104 1.175, - 250/2500 3155 62.5 1.104 the wave shape, it may be better to simulate the impulse generator as a capacitance and resistance network, as shown in Fig. 7.9 for a simple single-stage impulse generator. The initial voltage across C, would be nonzero, and the switch closing would simulate the gap firing. Fig. 7.10 compares measurements against EMIP simulation results for the waveshape of a multistage impulse generator, where the generator was modelled as a network of capacitances, resistances and inductances [127]. The spark gaps were represented as time-dependent resistances based on Toepler's formula. close at close at to R = 20 \-\- T object wor~o AE 2 (0) ¥0 T cy + S (a) Chreutt type a (>) Cireute type b Fig. 7.9 - Single-stage impulse generators 7-9 | | ; | 7 | i = \ soe ! 7 i] a 1 ° 18 io (a) Measurement (b) Simulation results (1 = exact with Ronsimultaneous firing of spark gaps, 2 = simultaneous firing, 3 = simultaneous firing with gaps as ideal switches) Fig. 7.10 - Waveshape of a multistage impulse generator [127]. © 1971 IEEE 7.6 Gurrent-Controlled de Voltage Source This source provides a simplified model of an HVDC converter station [128], and produces simulation results which come reasonably close to field tests [129]. The current-dependent voltage source is connected between two nodes (cathode and anode), as indicated in Fig. 7-11. The current can only flow in one direction (from anode to cathode). This is simulated internally with a switch on the anode side, which opens to prevent the current from going negative and closes again at the proper voltage polarity. Spurious voltage oscillations may occur between the anode and cathode side after the switch opens, unless the damping circuits across the valves are also modelled. Good results were obtained in [128] when an RC branch was added between the anode and cathode (R = 9000 and C= 0.15 us in that case). The current regulator is aseuned to be an amplifier with two inputs (one proportional to current bias Ig;,cs and the other proportional to measured current 1), and with one output ¢, which determines the firing angle. The transfer function of the regulator is 7-10 K(148T,) G(s) = : 7. (8) = Cat yd) (7-8) with Limite placed on the output e, in accordance with rectifier minimum firing angle, or inverter minimum extinction angle. The current-controlled de voltage source is a function of ¢,, Vac TH te Sy > (7-9) as shown in Fig. 7.12. The current regulator output e,, minus a bias value (1ov in Fig. 7-12) 4s proportional to cost. The taverter noraally operates at minimum extinction angle at the limit e,,,,, and the rectifier normally operates on constant current control between the limits. The user defines steady-state Limits for vs. which are converted to limits one, with Eq. (7.9). I the converter operates at the maximum limit e,.,, (or at the minimum Limtt eg¢q)> transient simulation, it will back off the limit as soon as the derivative de, /dt becomes negative (or positive) in the differential equation either in initial steady state or later during the dey (1, +) AE = KO gag) - km HE- hs (7.10) ‘The value for d?e,/dt? is zero in Eq. (7-10) as long as the converter operates at the limit. node name for cathode side 1a node nane for anode side Fig. 7.11 - Current-controlled de voltage source wu 150 Vac max Vag RY) oj 0 20 a -150 or -1l0 3 1.0 cos a Fig. 7-12 - Relationship between vg, and e, (ky = -150 000, ky = 15 000) 726-1 Steady-state dc initial conditions are automatically computed by the program with the specified value vq (0). Since the steady-state subroutine was only written for ac phasor solutions, the de voltage is actually represented as v4, = v4, (0) cos(wt) with a very low frequency of f = 0.001 Hz. Practice has shown that this is sufficiently close to dc, and still makes reactances wL and susceptances wC large enough to avoid numerical problems in the ac steady: tate solution. When the current-controlled de voltage source was added to the EMTP, voltage sources between two nodes were not yet permitted. For the steady-state solution, a resistance Rasy ts therefore connected in series with the voltage source, which is then converted into @ current source in parallel with R,,,+ This produces accurate results if the user already knows what the initial current 14,(0) is, because the specified voltage source of the rectifier ts automatically sed 4, (0), and that o averter is dec imereased by Roi, 1y,(0)» that of the inverter 1s decreased by Rowiy 7-2 14.(0)- The progran user should check, however, whether the computed current 1g, does indeed agree with what the user thought 1t would be, ‘This nufsance of having to specify 1,,(0), without knowing whether it will agree with the computed value, could be removed by using the methods described in Section 6.3, 1£ this HVDC model is used often enough to warrant the program changes The value of Rguty (Section 7.5.2). ‘The normal steady-state operation of an HVDC transmission link, measured is the same as the one used in the transient solution somevhere at a common point (e-g-, in the middle of the line) is indicated in Fig. 7-13. For the converter operating between the limits on constant current control (which is normally the rectifier), I,,,, 18 automatically BIAS computed to produce the characteristic A-A' of Fig. 7.13, (0) Tyrag 7 100) +e — > SE Cg < ey < Soman * qu) with 1(0), e,(0) being the de initial conditions. For the converter operating at maximum or minimum voltage (which is normally the inverter), the current setting I, must be given as part of the input, which defines SETTING the point where the converter backs off the limit and goes into constant current control. Ty,,, 18 again automatically computed, which in this case is e,00) Ty1as * Tserrinc * —K— tf ¢g(9) * ymax °F Cumin * (7-12) Iserrine £8 typically 15% lower than the current order Ipangq at the steady-state operating point for inverters (or 15% higher for rectifiers). 7.6.2 Trangient Solution In the transient solution, the dynamics of the current controller in the form of Eq. (7-9) and (7-10) must obviously be taken into account. First, rewrite the second-order differential equation (7.10) as two first-order differential equations, 7-13 HHP Em Ky rgg - 1) ~ RE a (7-13a) de x-qts (7.130) with the new variable x and with the new paraneters T= +h, (7.13) Pett (7a3ay After applying the trapezoidal rule of integration to Eq. (7-13a) and (7-13b) (replacing x by [x(e-te) + x(t)]/2 and dx/at by [x(t) - x(t-Ae)] /Ot, etc.), ‘and after eliminating x(t), one linear algebraic equation between e,(t) and 4(t) ds obtained. Inserting this into Eq. (7.9) produces an equation of the form v, = volt) - 1s ac(®) * Volt) ~ Beguayt(t) » (Talay which is a simple voltage source vo(t) in series with an internal resistance Reguty’ Tas Thevenin equivalent etreuit te converted {nto @ current source to(t) in parallel with Reougy (Pig- 7-14). rectifier tnvert rece eral yevRecT -Viny 4 \ inverter Tsertmg~ | Tonner 7 Pig. 7.13 - Normal operation of HVDC transmission link 7-14 » CATHODE R ‘equiv Ne aoe BOTTOM Fig. 7.14 - Norton equivalent circuit The equivalent resistance R,,,, remains constant for a given step size At, 2r 21) a RM BS FL de (at)? whereas the current source {o(t) depends on the values e,(t-At) and x(t-At) (7.15) of the preceding time step. After the complete network solution at each time step, with the converter representation of Fig. 7-14, the current is calculated with Eq. (7-14), and then used to update the variables e, and x. Ife, hits one of the limits ©, 45, OF Cain appropriate limit in the following time steps, with x and dx/dt set to zero. » it is kept at the B.C. Chiu has recently shown, however, that simply setting x and dx/dt to zero at the limit does not represent the true behaviour of the current controller [130]. The treatment of limits should therefore be revised, if this current-controlled de voltage source remains in use. Backing off the Limit occurs when the derivative de, /dt calculated from Eq. (7-10) becomes negative in case of e, = €,,,,+ OF positive in case of e, = ‘amax’ Samin* The switch opens as soon as i(t) < 0, and closes again as soon as Vanope > Yporrom’ t® assure that current can only flow in one direction. This updating of the current source io(t) from step to step ts not influenced by the switching actions. &1 8. THREE-PHASE SYNCHRONOUS MACHINE Co-author: V. Brandwajn The details with which synchronous machines mst be modelled depend very much on the type of transient study. Most readers will be familiar with the simple representation of the synchronous machine as a voltage source E” behind a subtransient reactance X4 short-circuit studies with steady-state phasor solutions, and is also + This representation 1s commonly used in reasonably accurate for transient studies for the first few cycles of a transient disturbance. Switching surge studies fall into that category. Another well-known representation is E' behind X,' for simplified stability studies. Both of these representations can be derived from the same detailed model by making certain assumptions, such as neglecting flux linkage changes in the field structure circuits for E" behind X,", and in addition, assuming that the damper winding currents have died out for E' behind X,'. The need for the detailed model described here arose in connection with subsynchronous resonance studies in the mid-1970's. In such studies, the time span is too long to allow the use of simplified models. Furthermore, the torsional dynamics of the shaft with ite generator rotor and turbine rotor masses had to be represented as well. Detailed models are now also used for other types of studies (e.g-, simulation of out-of-step synchroni- zation). To cover all possible cases, the synchronous machine model repre~ sents the details of the electrical part of the generator as well as the mechanical part of the generator and turbine. For studies in which speed variations and torsional vibrations can be ignored, an option is provided for by-passing the mechanical part in UBC's EMTP version MicroTran’. ‘The synchronous machine model was developed for the usual design with three-phase ac armature windings on the stator and a de field winding with one or more pole pairs on the rotor. For a reversed design (armature windings on the rotor and field winding on the stator), it is probably possible to represent the machine in some equivalent way as a machine with 8-2 the usual design. Even though the model was developed with turbine-driven generators in mind, it can be used for synchronous motors as well (e.g-, pumping mode in a pumped-storage plant). The model cannot be used for dual-excited machines (one field winding in direct axis and another field winding in quadrature axis) at this time, though it would be fairly easy to change the program to allow for it. Since such machines have not yet found practical acceptance, it was felt that the extra programming was not justified. Induction machines can also not be modelled with it, though program changes could again be made for that pur- pose. For these and other types of machines, the universal machine of Section 9 should be used. While the equations for the detailed machine model have been more or less the same in all attempts to incorporate them into the EMTP, there have been noticeable differences in how their solution is interfaced with the rest of the network. K. Carlsen, E-H. Lenfest and J.J. LaForest were probably the first to add a machine model to the EMTP, but the resulting "MANTRAP” - pro- gram [97] was not made available to users outside General Electric Co. M.C. Hall, J. Alms (Southern California Edison Co.) and G. Gross (Pacific Gas & Electric Co-), with the assistance of W.S. Meyer (Bonneville Power Administration), implemented the first model which became available to the general public. They opted for an iterative solution at each time step, with the rest of the system, as seen from the machine terminals, represented by a three-phase Thevenin equivalent circuit [98]. To keep this “compensation” approach efficient, machines had to be separated by distributed-parameter lines from each other. If that separation did not exist in reality, short artificial “stub lines” had to be introduced which sometimes caused problems. V. Brandwajn suggested another alternative in which the machine is basically represented as an internal voltage source behind some impedance. The voltage source is recomputed for each time step, and the impedance becomes part of the nodal conductance matrix [G] in Eq- (1.8). This approach depends on the prediction of some variables, which are not corrected at one and the same time step in order to keep the algorithm non-iterative. While the prediction can theoretically cause numerical instability, it has been refined to such an extent by now that the method has become quite stable and reliable. Whether 83 an option for repeat solutions as correctors will be added some day remains to be seen. Numerical stability has been more of a problem with machine models partly because the typical time span of a few cycles in switching surge studies has grown to a few seconds in machine transient studies, with the step size At being only slightly larger, if at all, in the latter case. 8.1 Basic Equations for Electrical Part Since there is no uniformity on sign conventions in the Literature, the sign conventions used here shall first be summarized: (a) ‘The flux Linkage A of a winding, produced by current in the same winding, is considered to have the same sign as the current (A=Li, with L being the self inductance of the winding). (b) The “generator convention” is used for all windings, that is, each winding k 1s described by aa Ct) yj (0) = RACE) - @-) (with the “load convention”, the signa would be positive on the right-hand side). (c) The newly recommended position of the quadrature axis lagging 90° behind the direct axis in the machine phasor diagram is adopted here [99]. In Park's original work, and in most papers and books, it is leading, and as a consequence the terms in the second row of [T]7} of Eq. (8-7b) have negative signs there. The machine paraneters are influenced by the type of construction. Salient-pole machines are used in hydro plants, with 2 or more (up to 50) pole pairs. The magnetic properties of a salient-pole machine along the axis of symmetry of a field pole (direct axis) and along the axis of symmetry midway between two field poles (quadrature axis) are noticeably different because a large part of the path in the latter case is in air (Fig. 8-1a). Cylindrical-rotor machines have long cylindrical rotors with slots in which distributed field windings are placed (Fig. 8.1b). They are used in thermal plants, and have 1 or 2 pole pairs. For cylindrical-rotor machines the mag- a4 netic properties on the two axes differ only slightly (because of the field windings embedded in the slots), and this difference ("saliency") can often be ignored, especially for unsaturated conditions where X3,°K,,- The word saliency is used as a short expression for the fact that the rotor has different magnetic properties on the two axes. The machine model in the neck (a) Salient-pole machine (b) Cylindrical-rotor machine Fig. 8.1 - Cross sections of synchronous machines (d = direct axis, q = quadrature axis) [101]. Reprinted by permission of I. Kimbark EMTP always allows for saliency; if saliency is ignored, the same equations will still be used, except that certain parameters will have been set equal at input time (KX, = Ky, etc.). The electrical part of the synchronous machine is modelled on a per phase basis with 7 coupled windings”): “Yanother, more modern approach is to measure the frequency response from the terminals, which can then be used to represent the machine with transfer functions between the terminals, without assuming a given number of lumped windings a priori. One can also use curve fitting techniques to match this measured response with that from a series and parallel combination of R-L branches [100]. The end result in the latter case is basically the same model as described here, except that damper bars are sometimes represented by more than one winding, and that the data is obtained from frequency response tests. 8-5 three armature windings (connected to the power system), one field winding which produces flux in the direct axis (connected to the dc source of the excitation system), one hypothetical winding in the quadrature axis to represent slowly changing fluxes in the quadrature axis which are produced by eddy currents in the solid rotor of cylindrical-rotor machines (effect normally negligible in salient-pole machines), one hypothetical winding in the direct axis to represent damper bar effects, one hypothetical winding in the quadrature axis to represent damper bar effects, or the faster changing fluxes produced by eddy currents in the solid rotor of cylindrical-rotor machines. For machines with more than one pole-pair, the electrical equations are the same as for one pole-pair, since modelling is performed on a per phase basis, except that the angular frequency and the torque being used in the equations of the mechanical part must be converted as follows: ¥ wel “gech ~ p72 (8.28) = Tech * Ten (8.2b) where p/2 is the number of pole pairs. The behaviour of the 7 windings is described by two systems of equations, namely by the voltage equations y-4 tv) = -RItG- & 1, (8.3) with Gl = Gy, igs dg, igs dg, tp, agl, TAI = Days Ags Age Age Age Ape Aad (v] = [¥), v9» vgs Ver 0, 0, 0] (zero in last 3 components because g-, D- and Q- windings are short-circuited), (R] = diagonal matrix of winding resistances R,, R,, Ry» Res Rey Rp, Rg (subscript "a" for armature), and by the flux-current relationship 8-6 OQ) = (L1G) with (1) = | 1, (8.4) To make the equations manageable, a number of idealized characteristics are assumed, which are reasonable for system studies. These assumptions for the “ideal synchronous machine" [76, p. 700] are *): q@ @) @) @ (5) (6) >) (8) (9) *) ‘The resistance of each winding is constant. ‘The permeance of each portion of the magnetic circuit is constant (corrections for saturation effects will be introduced later, however) . ‘The armature windings are symmetrical with respect to each other. The electric and magnetic circuits of the field structure are symmetrical about the direct or quadrature axis. The self inductance of each winding on the field structure (£,g,D,Q) is constant. The self and mutual inductances of the armature windings are a constant plus a second-harmonic sinusoidal function of the rotor position B (second-harmonic component zero if saliency ignored), with the amplitude of the second-harmonic component being the same for all self and mutual inductances. The mutual inductance between any winding on the field structure and any armature winding is a fundamental sinusoidal function of the rotor position B. Effects of hysteresis are negligible. Effects of eddy currents are negligible or, in the case of cylindrical-rotor machines, are represented by the g-winding. For a detailed design analysis of synchronous machines, many of these idealizations cannot be made, Since they imply that the field distribution across a pole is a fundamental sinusoid, spatial harmonics produced by the nonsinusoidal field distribution in a real machine could not be studied with the ideal machine implemented in the EMTP. 8-7 ‘Then, Ly, 7 Lt 1, cos28, similar for Ly), Ly, Lyy * Lay = My + 1, cos(2B-120°) similar for Lj, Ly = Myg COsB, similar for Lye. Ly- = May cosB, similar for Ly, Ly) (8.5) Lig gi "Mag 3108. similar for Ly, byg Lig = Hg, = Mag SinB, similar for Ly, lyg Leer Lag, Epps Lage Mens Mgg constant (not functions of B) , with B being the angular position of the rotor relative to the stator (8... h = a? which is related to the angular frequency, p ag oe. (8.6) Sone authors (e.g., Kimbark [101]) use a different sign for M, in Eq. (8.5). With the sign used here, the numerical value of M, will be negative. The solution of the two systems of equations (8.3) and (8.4) is compli- cated by the fact that the inductances in Eq, (8.4) are functions of time through their dependence on B in Bq. (8.5). While it is possible to solve them directly in phase quantities ") from phase quantities to d, q, 0- quantities because the inductances become + most authors prefer to transform them constants in the latter reference frame. This transformation projects the rotating fluxes onto the rotating field axis, from where they appear as stationary during steady-state operation. The transformation was first proposed by Blondel, and further developed by Doherty, Nickle and Park; in North America and in Europe, it is now often called Park's transformation. The transformation is identical for fluxes, voltages, and currents, and converts phase quantities 1,2,3 into d, q, 0- quantities, with quantities on the field structure ri ining unchanged, “1g space harmonics in the magnetic field distribution had to be taken into account, then L1, and Ly) in Bq. (8.5) would have added 4th, 6th, and higher harmonics terms, and Ly, etc. would have added 3rd, 5th,...harmonics terms. In that case, solutions in phase quantities would probably be the best choice. Mago) = (1-4 DAL, identical for (v1, 4] , (8.78) ith Dayo) “igs Agr Nor Mer Mgr Mar Mgl and remain unchanged B cos p 12 cos (p-120°) 12 cos (pri20°) 0000 a a a 2 sing 12 sin (p-120°) ¥2 sin (pt120°) 0000 a a a (r] = — — 0000] (8.7b) a a a ° ° ° 1000 ° o o 0100 ° o ° 0010 0 ° ° 0001 Bq. (8.7) is an orthogonal transformations it therefore follows that 1] = [T)-f,ensposed (8.8) ‘The matrices [T] and [T]-1 are normalized here. This has the advantage that the power is invariant under transformation, and more importantly, that the inductance matrix in d, q, 0- quantities is always symmetric. The lack of symmetry with unnormalized quantities can easily lead to confusion, because it is often removed by rescaling of field structure quantities which in turn imposes unnecessary restrictions on the choice of base values if p.u. quantities are used. Authors who work with unnormalized transformations use a factor 2/3 in the first two rows of Eq. (8.7b), and 1/3 in the third row. In many older publications the position of the quadrature axis is assumed 90° ahead of the direct axis, rather than lagging 90° behind as here, and the second row of Eq. (8.7b) has therefore negative signs there. Transforming Eq. (8.3) to d, q, 0- quantities yields (ago) = “RV Sago) - de Pago) + (8.9) a9 which is almost identical in form to Eq. (8.3), except for the "speed voltage terms" anal and wr (voltages induced in armature because of rotating field poles). They come out of Eq. (8.3) by keeping in mind that [T] is a function of time (B = ut), tr)-+ 207) faggg}) = gf ggg) + (1-4! 71) Digg) + a ‘dqo at ago ‘ae ‘ago ‘Teansforming Eq. (8.4) yields flux-current relationships which can be partitioned into two systems of equations for the direct and quadrature axis, and one equation for zero sequence, ta Mar Map a de || Mar Mee Meo (8.108) Mey Ly dot, Mg Ly Mag Mag neal 8, 10b) Mae ‘ee “eo ae re Fa “ea Ma 3 8 whe: By gt By» where Mag 5 Mag’ Mag ~ V5 Maa and Ay * Loi (8.10¢) Most elements of these inductance matrices with constant coefficients have already been defined in Eq. (8.5), except for direct axis synchronous inductance quadr. axis synchronous inductance rere) c Ly (8.11) zero sequence inductance 8-10 8.2 Determination of Electrical Parameters” A set of resistances and of self and mutual inductances is needed in the two systems of equations (8.9) and (8.10), which are not directly available from calculations or measurements, According to IEEE or IEC standards [102,- 103] the known quantities are armature resistance R, armature leakage reactance % 3 zero sequence reactance Xs synchronous reactances yr Xy transient reactances X's Xl subtransient reactances XX" transient short-circuit time constants 74’, Ty’ subtransient short-circuit time constants T,", T)" All reactances and time constants must be unsaturated values, because satura~ tion is considered separately, as explained in Section 8.6. This is the reason why short-circuit time constants are preferred as test data over open- circuit time constants, because the measurement of the latter is influenced by saturation effects [104]. One set of time constants can be converted into the other set if R, is ignored. In [104] the conversion is done with Xq Tas + Tas" xr Tal + x a WT" Kyra (8,12a) x Nola * TaN! & a for the direct axis, and identically for the quadrature axis by replacing subscript "d" with "q". This conversion has a very small error, which is negligible for typical machine data. For a precise conversion, still assuming R, = 0, Eq. (8.12a) is replaced by WTyg') (T, Ty! Ty 1 a ry, (8.12) a Tao) Ty" Tag") Tj" GT) T+ wi? “the assistance of S. Bhattacharya and Ye Zhong-liang in research for this section is gratefully acknowledged. as explained in Appendix VI.3. Eq. (8.12b) becomes Eq. (8.12a) as the factors (wT)?/(1+(wT)?) approach the value 1.0. The number of known parameters is less than the number of resistance and inductance values in Eq. (8.9) and (8.10), and some assumptions must there- fore be made before the data can be converted. Since the procedure for data conversion is the same for the direct and quadrature axis parameters, only the direct axis will be discussed from here on. Winding D is a hypothetical winding which represents the effects of the damper bar squirrel cage. We can therefore assume any number of turns for it, without loss of generality. In particular, we can choose the number of turns in such a way that Man . Mag (8,13a) in Eq. (8.10a), and similarly Mag ~ “og in Eq. (8,10b), Many authors represent the flux-current relationships with an equivalent star circuit, which requires all three mutual inductances in Eq. (8,10a) to be equal. This is achieved by modifying (rescaling) the field structure quantities as follows: Dem = 2 en and 4, oa (8.14a) ‘fm fm t 2 Mat with k= 5 (8.14) £0 = = 4 ‘a in n a dem |] Me Uetm Me is (8.152) Dom ee AOD Som with . 3 ya = 3 xg and Rem = 3 K*Res Roy * 5 KAR (8.16) 8-12 Fig. 8.2 - Equivalent circuit for direct axis with modified field structure quantities Fig. 8.2 shows the equivalent star-circuit for the direct axis, with the speed voltage term and resistances added to make it correct for Eq. (8.9) as well. Modifying the field structure quantities is the same as changing the number of turns in the field structure windings. It does not influence the data conversion, but it is simpler to carry out if the modified form of Eq. (8.15a) is used. The correct turns ratio is then introduced later from the relationship between rated no-load excitation current and rated terminal voltage. The best data conversion procedure seems to be that of Canay [104]. It uses the four equations which define the open- and short-circuit time constants, as derived in Appendix VI.2, to find Re, Ryas Lee, and Lp, ("n" dropped in Appendix VI to simplify the notation). Usually, only one pair of time constants (either Ty}, Ty” or Ty’, 74") plus X,', X," are known; in that case, the other pair must first be found from Eq. (8.12). Solving the four equations for the four field structure quantities presupposes that the mutual inductance M, in Bq. (8.154) i already known, Its value has traditionally been found from leakage flux calculations. While turns ratios have been unimportant so far, they mist be considered in the definition of leakage 8-13 flux, since it is the actual flux 9, rather than the flux linkage A = Np (N = number of turns) which is involved. In defining the leakage flux we must either use actual flux quantities, or flux linkages with turns ratios of The leakage flux Linkage produced by 1, is then Ag 7 bgtge Lggtg » Provided NyiN, = 1:1 (8-17) Let us assume that all field structure quantities are referred to the amature side, which implies N,:N, = 1:1 in the original equations (8.4) with phase quantities, with the mtual inductance being M,- (cos p = 1.0 if magnetic axis of phase 1 armature winding lined up with direct axis). After transforming to 4, q, 0- quantities, the mtual inductance in Bq. (8.10a) By, ‘a 3 = "311. To convert back to a ratio of 1:1, the second row and column in rea Eq.(8-10a) mst be multiplied with “2 , which changes the mtual inductance wz co 2 Myge Ten the leakage flux Linkage produced by ty wth a tel catto vecones = oa Re “bata 2 Macha > or for the Leakage inductance, Ly tly - Unfortunately, this equation is still not enough for finding M, in the (8.18) modified matrix of Eq. (8.15a) because of the unknown factor k in Eq. (8-14b). To find k, an additional test quantity must be measured which has not yet been prescribed in the IERE or IEC standards. It has therefore been common practice to assume k = 1, which implies Mp) = Mag» With this assumption, the results for armature quantities will be correct, but the amplitude of the fast oscillations in the field current will be incorrect, as pointed out by Canay [104] and others. Fig. 8.3 shows the measured field current after a three-phase short-circuit [104], compared with EMTP simulation results with k = 1, and with the correct value of k. Note that B14 4.0 2.0 ceed == correct k field test 0.0 0.0 0.05 0.10 0.15 0.20 —t (s) Fig. 8.3 - Field current after three-phase short-circuit [104]. Reprinted by permission of IEE and the author “branch” in the star circuit of Fig. 8.2 is the leakage inductance only 41£ all quantities are referred to the armature side and if k= 1. If the factor k is known, then the “d-branch” with field structure quantities referred to the armature becomes Canay's "characteristic inductance” 3 Ly thy - Zk he (8-19) The data conversion of the modified quantities on the direct axis can tow be done as follows: If k is unknown, assume k = 1, find M, from Eq. (8-18), a (8-20a) and realize that the fast oscillations in the field current will have a wrong amplitude, but the armature quantities will be correct. If the characteristic inductance is known (which can be calculated from k), find M, from Eq. (8.19), in“ Ya and the fast oscillations in the field current will be more correct. Then use the conversion procedure of Appendix VI.4 to obtain the field structure quantities Roos Rog» Leggs Lppg ("m" dropped in Appendix VI), which will be rescaled according to Eq. (8.14). It is not necessary to undo the rescaling “bye (8.20b) if one is only interested in quantities on the armature side, because scaling of field structure quantities does not influence the armature quantities. If the conversion was done with p.u. quantities, which will usually be the case, then multiply all resistances and inductances with V3,..4/S, rated’ Srated physical values (V, rated line-to-line RMS armature voltage, S. to obtain rated ‘rated ~ rated apparent power) for wye-connected machines, followed by another multiplication with a factor 3 for delta-connected machines. The data conversion for the quadrature axis quantities is the same as that for the direct axis, except that one does not have to worry about correct amplitudes in the oscillations of the current i,. This current cannot be measured, because the g-winding is a hypothetical winding which represents eddy or damper bar currents. It is therefore best to use k=l end (8.20¢) Ma (quadrature axis) on the quadrature axis. Rather than undo the rescaling of Eq. (8.14) by using t-1/( 43 &) with the procedure described after Eq. (8.22), it makes more sense to eee factor t which introduces the correct turns ratio between physical values of the armature and field structure quantities. To find this factor, we must look at the open-circuit terminal voltage produced by the no-load excitation current 4, For open circuit, ae 0, and, in steady-state 8-16 operation, dd,/dt = 0, which leads to Vg 7 My i Since we know that the modified current must be t-times the actual current, fm Leg = tf, f and since v, is equal to 13+V, nyc (assuning symmetrical voltages in the three phases), we can find t from 3 V hase ee (8.21) WM 2 eno load with Vphase = rated RMS line-to-ground voltage for wye connection, and line-to-line voltage for delta connection, rated no-load excitation current which produces rated voltage “no load ~ at the terminal. Sometimes the no-load excitation current is not known. Then any system of units can be used for the field structure. One possibility is to set t = 1 (field structure quantities referred to the armature side). Another possibility is to say that a field voltage |v-| = 1.0 should produce the rated terminal voltage. Then 7 Wel ao f-no load ~ R, ae ce which, when inserted into Eq. (8.21) gives wt, t-——2—_.. (8.22) 3 Vonase Bt once t is known, the inductances are converted by multiplying the second and third row and colum of the inductance matrix in Eq. (8.15a) with t, and by multiplying R,, and Ry, with t?, The quadrature axis inductances and resistances are also multiplied with t or tt, respectively. Sometimes, generators are modelled with less than 4 windings on the field structure (D-winding on the direct axis missing, and/or g- or Q- winding on the quadrature axis missing), In such cases, the EMIP still uses the full 7-winding model and simply "disconnects" the unvanted winding by setting its off-diagonal elements in the inductance matrix to zero, and its 8-17 diagonal element to an arbitrary value of ul = 10. Its resistance is arbitrarily set to zero, The inductances and resistances of the other windings on the same axis are calculated from Eq. (VI.4) and (VI.5) (Appendix VI), e.g., for a missing g-winding, - =0,R = 0, ub =1, Mag ~ 9 Mag = OF Rg 7 Or Wlyg = 2 and + Tag = Rg Tyo" > where My = 1) - L 90 Caan With one winding missing, the characteristic inductance L, does not come into the picture, and the leakage inductance L, must always be used in calculating M,» on the direct as well as on the quadrature axis. 8.3 Basic Equations for Mechanical Part There are many transient cases where the speed variation of the generator is so small that the mechanical part can be ignored. Simulating short-circuit currents for a few cycles falls into that category. In that case, w in Eq, (8.6) and in the other equations is constant, and the angular position B of the rotor needed in Eq. (8.7) and (8.9) is simply B(t) = B(O) + ut, (8.23) with B and w being angle and speed on the electrical side. For other types of studies it may be necessary to take the speed variations into account. The simplest model for the mechanical part is the single mass representation as used in stability studies, 2B y B - J, + de ~ Teurbine ~ Tgen * (8.24a) ae and Baw (8,24b) with J = moment of inertia of rotating turbine-generator mass, B = rotor position, . = speed D = damping coefficient for viscous and windage friction (linear dependence on speed is crude approximation) , Trurbine 7 torque input to turbine, Tyen = electromagnetic torque of generator. Eq. (8.24) is valid for quantities referred to the electrical or the 8-18 mechanical side with the conversion from one to the other being *) Tea = Tmech/ (P/2)* Bee = (P/2) Brech (8.25) Dee = Pmech/ (P/2)? Tee = Tnecn/(P/2) With voltages given in V, and power in W, the unit for the torque T becomes Nm, for the damping coefficient D it becomes Nm/ (rad/s) and for the moment of inertia J it becomes kg-m? (kg as mass). In the English literature, the moment of inertia is sometimes given in slug-ft?; then 1 slug-ft? = 1.355818 kg-m*. The moment of inertia may also be given as a quantity WR? (concentrated weight of mass W rotating at radius R). With J = wR’/g in slug-ft? and 932.174 ft/s?, we get 1 WR? (1bgaggft?) = 0.0421401 keg-m?. Instead of the moment of inertia, the kinetic energy at synchronous speed is often given, which is identical for the mechanical and electrical side, E=% Inech noch = ¥ Jer Oop (8.26) The inertia constant h (in seconds) is the kinetic energy E (e.g. in kWs) divided by the generator rating Srating (€-9-, in kVA), E ating he (8.27) The relationship between the inertia constant h and the acceleration time T, of the turbine-generator is Ty Pra he spy iutine (8.28) with Frating = Fated power of turbine-generator (e.g., in MW), Srating = rated apparent power of generator (¢.g., in MVA). A single mass representation is usually adequate for hydro units, where turbine and generator are close together on a stiff shaft. It is not good » Subscript "mech" is used for the mechanical side, and subscript "ee" for the electrical side. 8-19 enough, however, for thermal units, if subsynchronous resonance or similar problems involving torsional vibrations are being studied. In such cases, a number of lumped masses mst be represented. Usually 6 to 20 lumped masses provide an adequate model”, ‘The model in the EMIP allows any number of lumped masses n > 1, and automatically includes the special case of n= 1 in Bq. (8.24). Each major element (generator, high pressure turbine, etc.) is considered to be a rigid mass connected to adjacent elements by massless springs. Fig. 8.4 shows a typical 7-mass model. The shaft/rotor system is assumed to be linear, which is reasonable for the small amplitudes of typical torsional vibrations. The n spring-connected rotating masses are then described by the rotational form of Newton's second law, fcurbine 1 Teurbine 2 Teurbine 2 Teurbine 4 Tturbine 5 uP Fy A Py pe ex exc Fig. 8.4 Mechanical part of a steam turbine generator with 7 masses (HP = high pressure turbine, IP = intermediate pressure turbine, LPA, LPB, LPC = low pressure turbine stages A, B, C, GEN = generator, EXC = exciter) (5) (0) + 0) gf (ol + (4) (0) = Myuentne) ~ Fyenvere! 0299 and 2) = wl (8.290) where *\mere are studies where the lumped mass representation is no longer adequate, and where continuum models must be used. 8-20 [3] = diagonal matrix of moments of inertia (J,,.+.J, in Fig. 8-4), i) = vector of angular positions (@,,.-.0, in Fig. 8.4, with 0, = BD» lw) = vector of speeds, (] = tridiagonal matrix of damping coefficients, {x] = tridiagonal matrix of stiffness coefficients, (Tesepine! ~ Vector of Eorques applied to the turbine stages (Teurpineg "0 and Tyuepiney 7 0 1m Fig. 8-4), [Tyensexe) 7 Vector of electromagnetic torques of generator and exciter (components 1 to 5 = 0 in Fig. 8.4). The moments of inertia and the stiffness coefficients are normally available from design data. The spring action of the shaft section between Oy > constant K es i-l and i creates a torque which is proportional to the angle twist The proportionality factor is the stiffness coefficient or spring ‘t-1,1° Ts spring action torque acts in opposite directions on masses i-1 and 1, Tepring,t-1 ~~ Tepring,t ~ Xi-1,1(1-17 4) + (8.30) If these torques are included in Eq. (8.29), they create the term (K][0]. Prom Bq. (8-30) it can be seen that [K] has the following form 12 Kz “Kin Riz + Koy Kay (x) = Kp Kay * Koy “Koy *yijn Kiyo Two damping effects are included with the damping coefficients, namely the self-danping D, of mass i (mostly friction between the turbine blades and stean), and the danping Dy, , created in the shaft when the shaft between masses 1-1 and { is twisted with the speed w,- wy; The damping torque acting on mass 1 1s therefore 1, «>, hep, (os -— yep (Bt - Pity (8.31 ‘damping, ~ a “ae * Paya (ae 7 ae) + aa Cae mae) OSD) From Eq. (8-31) it can be seen that [D] has the same structure as [K], except 8-21 that the diagonal element is now D,_ D,+D,, -D. (D] = win Pn-tyn * Pp It is very difficult to obtain realistic values for these damping coefficients. Fortunately, they have very little influence on the peak torque values during transient disturbances, However, for estimating the low cycle fatigue one must consider the damping terms, which, unfortunately until now, have often been derived from unsuitable models [107]. The damping of torsional oscillations is measured by observing the rate of decay when the shaft system is excited at one of its natural frequencies (modes). It is difficult to convert these modal damping coefficients into the diagonal and off-diagonal elements of [D]. For the vector of turbine torques it is best to assume that the turbine power P, = wT, ‘turbine ‘turbine decreases with an increase in speed, which is more reasonable than constant remains constant. This implies that the torque torque because if the turbine were to reach the same speed as the steam (or water) jet the torque on the blades would obviously drop to zero. It is possible to include the effects of the speed governor through TACS modelling, but usually the time span of transient simulations is so short that the governor effects will normally not show up. The vector of electromagnetic torques and the rotor position of the generator provide the link between the equations of the mechanical and electrical part, with ee Snech-gen * 2 ~ Beg * (8.328) -2 eo Trech-gen ~ ~2” aig ~ Agia) » (8.32b) i, + igh, t * 1 Foxe Tnech-exe (8.32¢) “mech where it is assumed that (Eq. 8.29) is written for the mechanical side. If 8-22 it is written for the electrical side, then the conversion of Eq. (8-25) must e term 1,2 be used. The term 1,28, the exeiter machine; the negative sign for v_ comes from the source in Eq. (8.32c) represents the losses incurred in convention of Eq. (8.1). If there is no exciter machine, as in the case of rectifier excitation systems, then mass no. 7 in Fig. 8.4 would obviously be deleted, and Eq. (8-32c) would not be needed. Toy 949 18 not in the BPA EMTP. 8.4 Steady-State Representation and Initial Conditions Transient studies with detailed turbine-generator models practically always start from power frequency steady-state conditions. The EMP goes therefore automatically into an ac steady-state solution whenever the data file contains turbine-generator models. In some versions of the BPA EMTP, the machine is represented as a symmetrical voltage source at its terminals. This approach is only correct 1£ the generator feeds into a balanced network. In that case, the generator currents are purely positive sequence. In an unbalanced network, there are negative and zero sequence currents, which would see the generator as a short-circuit. This is incorrect, because generators do have nonzero nega ‘zero’ BPA EMTP, the user can specify unsymmetrical voltage sources at the termi- tive and zero sequence impedances 2,,, and Z,.,,. In the 439 version of the nals. This is still not good enough, however, unless the user adjusts the negative and zero sequence components of the terminal voltage iteratively neg ~ *negtneg *™4 Yeero “ ~zerotzero’ With Thegr Tzero Detns the currents from the steady-state network solution. The UBC EMIP does include until V, negative and zero sequence impedances correctly, as explained next. The negative sequence impedance can be calculated as part of the data conversion. Its imaginary part is very close to (8-33) and this value can be used without too much error if the negative sequence impedance is needed for preliminary calculations. The real part Rig 1s larger than the armature resistance R, because of double-frequency 8-23 circulating currents in the field structure circuits; its value is difficult to guess, and is therefore best taken from the data conversion. For calculations internal to the UBC EMTP, the correct values from the data conversion are always used. The zero sequence impedance 2)... = R, + JKo.9 is part of the input data, but its value becomes immaterial if the generator step-up transformer is delta-connected on the generator side and if the disturbance occurs on the line side. The positive sequence representation can be a voltage source behind any impedance, as long as it produces the desired values for the terminal voltage Ypos and the current T.,,. Knowing only V,, may require a preliminary steady-state solution with a voltage source V,,, at the terminal, to find I,og* Then the value needed for the voltage source behind the impedance is v, ve Tee source ~ “pos * *Tpo8 Any value can be chosen for Z, but Z = 2... simplifies programming for the following reason: The EMIP solves the network in phase quantities, and assunes that all phase impedance matrices are symmetric. Only Z = Zygy will Produce a symmetric phase impedance matrix, however. Changing the program to handle unsynmetric matrices just for generators would have required a substantial anount of re~progranming. The generator positive sequence representation is then a voltage source behind 2... while the negative and zero sequence representations are passive tapedances Z,,, and Z,.,1 respectively (or zero voltage sources or short- ‘zero circuits behind 2, and Z,,,,)+ Converted to phase quantities, these 3 ‘zero single-phase sequence representations become a three-phase symmetrical voltage source (E,", By Fig. 8.5(a), with » Ej") behind a 3 x 3 impedance matrix, as shown in 7 2 a a a (2] =] 2, 2, 2] » where 1 (8-34) fa ta %s 7° 5 (zero ~ "neg) To be able to handle any type of connection, including delta and impedance- grounded or ungrounded wye connections, the voltage sources behind [2] are converted into current sources in parallel with (Z], as indicated in Pig. 8.5(b), with 8-26 (a) Thevenin equivalent circuit (b) Norton equivalent circuit Fig. 8.5 - Turbine-generator representation for steady-state solution [| -[ [E] (8.35) 13, BE” This is done because the EMTP could not handle voltage sources between nodes until recently and even after such voltage sources are allowed now, this Norton equivalent circuit is at least as efficient. For armature winding 1, the internal voltage source is yao +z, (8.36) pos * “neg "pos * With Viog> Tyo being the unnormalized positive sequence values. Unnorasiized values are a more convenient input form for the user, because with the unaormalized transformation of Eq. (6-38) the positive sequence values are identical with the phase values V, and I, for armature winding 1 for balanced network conditions. For armature windings 2 and 3, the tnteroal voltages are B," = a2E," and £5" = af. . 8.37) (a= eSt20 ) In the UBC EMTP, either V,,, and T,,, can be specified as input data, in which case E," 1s calculated from Eq. (8.36), or E," can be specified directly. Specifying Ej" is not unusual as it may seem, because 8-25, short-circuit programs use essentially the same generator representation (E” beind X,"). If users want to specify active and reactive power at the terminals, or active power and voltage magnitude, then the load flow option described in Section 12.2 can be invoked, which will automatically produce the required V,., and T,.,+ The UBC EMTP connects the generator model of Fig. 8.5(b) to the network for the steady-state solution, which will produce the terminal voltages and currents at fundamental frequency. For unbalanced network conditions, this solution method is not quite correct because it ignores all harmonics in the armature windings and in the network. Experience has shown, however, that such an approximate initialization is accurate enough for practical purposes. Fig. 8.6 shows simulation results for a generator feeding into a highly g 7 1:02, Ry = 0.059), with an initialization procedure which ignores the harmonics on the armature side, unbalanced load resistance (R, = 1.02, R and considers only the second harmonics on the field structure side, as discussed in Section 8.4.2. The final steady state is practically present from the start. The mechanical part is totally ignored in the steady-state solution, because it is assumed that the generator runs at synchronous speed. Again, for unbalanced conditions this is not quite correct because, in that case, the constant electromagnetic torque has oscillations superimposed on it i (ka) { 20 -20 ° 50 100 —> t (ns) Fig. 8.6 Steady-state behaviour of generator with highly unbalanced load 8-26 which produce torsional vibrations whose effects are ignored. After the steady-state solution at fundamental frequency has been obtained, the terminal voltages and currents are converted to unnormalized symmetrical components to initialize the machine variables: 7 o ‘zero A 1 Yoos |=] @ a2] |v, |, tdenticar for (8-38) . laa 7 currents, neg 3 (exe4!0"). the inverse transformation 1s 1 a zero v,}=]1 a2 a v » identical for currents (8-39) 2 ete 3 toa neg 8.4.1 Initialization with Positive Sequence Val Tg the positive sequence voltage is obtained as a peak (not RMS) phasor - dy alue |Viogle°'POS then from Eq. (8-39), W(t) = WVeogl 08 (ut + 1395) W208) * [Vjog | 608 (ut + 1955 ~ 120°) (8.40) vat) = [Vaogl ©08 (ut + Yy95 + 120°) Inserting these voltages into the transformation of Eq. (8.7) produces 3 a Yarpostt? * 72 aoe! $49 (Ypo5 ~ 8) ony Yeepos pos! ©8 (togg ~ 8) ee ce) = "= |v, v2 where 6 is the angle between the position of the quadrature axis and the real axis for phasor representations (Fig. 8.7). This angle is related to the rotor position B., on the electrical side by ek Bag ob HO +E (8.42) The positive sequence values vq and v, in Hq. (8-41) are de quantities and hence do not change as a function of time; the argument (t) can therefore be 8-27 dropped. From Eq. (8-41) it 1s evident that vz and v, can be combined into a complex expression ws + iy, ees pos * Ma-pos ~ > ‘pos ty. wi es ex quant Pos, 's a pha: th V4, being the complex quantity |¥,,,| Witte V,, 18 2 phasor of 6 (8.43) frequency w in the network solution reference frane, Vi, ©” in Eq. (8.43) becomes a dc quantity in the d, q-axes reference frame. Similarly, a topos * Hacpos "= 3 ‘a-pos * Ha-pos ~ 7 “pos ok (8.44) Je, POS with respect to the real with T,,, betng the complex current |), quadrature real axis for network phasor solution Fig. 8.7 - Definition of 6 axis. With V,, and I,,, known, we stil need the angle 6 to find the ,q-values. To calculate 6, use Eq. (8.9) and Bq. (8-10), with 4) = 4, = ig= ©, as well as ddg/dt = 0 and d,/dt = 0 (no currents in damper windings and all d,q-quantities constant in aymmetrical operation with positive sequence values only), 8-28 Va-pos ~ “Rata-pos ~ “Mgiq-pos Yerpos ” Falg-pos * Malapos * " Meets pos (2.45) Eq. (8.454) and (8.45b) can be rewritten as a complex equation relative to the quadrature axis (8.458) Ya-pos * J¥a-pos + Hggos) + Egpos (8-450) where Eo, is @ quantity whose position on the quadrature axis is important, but whose magnitude = R, + july ‘q-pos Farpos ~ (hg - gta pos * Marke-pos is immaterial here. By inserting Bq. (6.43) and Eq. (8.44) into Eq. (8.45¢), and by multiplying with (¥2/13)e)°, a11 de quantities become phasors in the network solution reference frame again. The angle 6 is then obtained from the phasor equation - Fapos #7” = Vpog + (Ry + ILI (8.46) (Fig. 8.8). quadrature axis 2 % RL ‘2 pos real axis for network phasor solution Fig. 8.8 - Calculation of angle 6 8-29 With 6 known, the initial values v, 0), v, 0), 4, 0), 4, 0 ‘a-pos(» “q-pos ©)» S4-pos 9)» tq-pos are found from Eq. (8-43) and (8.44). As mentioned before, the remaining currents 1), 1,5 tg are zero from the positive sequence effects, except for ig, whose initial value is calculated from Eq. (8.45b), Yq-pos * Batg-pos() “what = “acpos ‘ta: 16 p50) oe : (8.47) ae L¢pog(0) 18 used to intttalize vp Ye -pop( * Retepog( + (8.48) The initial value of the torque produced by the positive sequence quantities is needed in the mechanical part. It is calculated from the Fluxes Ag o9g() * Bala pog() + Maptp pog(O) aMd Re nog) * Kgtgapog(O) a8 (0) 0) 4. (0) = 2g p996 - a-pos ‘q-pos ‘q-pos, 44-pos sey) T gen-pos' The initial positive sequence torque can also be calculated from energy balance considerations (uf = power delivered to the network + losses in the armature windings), oe 2 7 2 Teen-pos ) * J ¥llpos¥pos! * Z |poe!® Ra Se (division by 2 because the phasors are peak values. 8.4.2 Initialization with Negative Sequence Values”) If the network 1s balanced in steady state, then there are no negative sequence values. This part of the initialization can therefore be skipped if the negative sequence (peak) phasor current Je, Te eee |e (8.51) neg ~ | Tneg obtained from the steady-state solution is negligibly small. Negative sequence currents in the armature windings create a magnetic field which rotates backwards at a relative speed of 2w with respect to the field structure. Second harmonic currents are therefore induced in all *) ‘The negative sequence, currents in the BPA EMTP can be incorrect (see beginning of Section 8.4). 8-30 windings on the field structure, which the EMTP takes into account in the initialization. These second harmonic currents induce third harmonics in the armature windings, which in turn produce fourth harmonics {a the field structure windings, etc. Fortunately, these higher harmonics decrease rapidly in magnitude. They are therefore ignored. Calculating the field structure harmonics of order higher than 2 could be done fairly eastly, but the calculation of the armature harmonics of order 3 and higher would involve solutions of the complete network at these higher frequencies. While network solutions for harmonics may be added to the EMIP some day, this addition does not appear to be justified for this particular purpose. First, the negative sequence current mst be defined on the direct and quadrature axis. By starting with the negative sequence currents in the three armature windings, 1102) * [Tygg! 608(Wt + a9) £208) = [Tyggl com(ut + aggg + 120°) (8.52) 15C2) © [Tagg! soa(ut + ayg9 - 120°) these d, q-axes values are obtained through the transformation (8.7), v3 Sacneg() "- & Mog! sin(angg +6 + 20t) (8-53) v3 Leeneg(t) * & Mugg! £9(dugg + 6 + 2at) While the positive sequence d, q-axes currents are de quantities, the negative sequence d, q-axes currents are second harmonics. This {s important to keep in sind when we represent then with a phasor of frequency 2u, Lageneg * 2 _ (8.54) with the understanding that Sot Lgeneg(t) 7 -Im(Ty ce ‘dq-neg eduty i-neg’ (8-55) te neg(t) * RefT, arneg ®) * Fe{Taq-neg 8-31, For the initialization of 1, and 1,, the negative sequence values at t+0 from Eq. (8.55) must be added to the respective positive sequence values from Eq. (8.44) to obtain the total initial values 14(0) and 1,(0). The negative sequence d, q-axes voltages are not needed in the initialization, but they could be obtained analogously to the currents. The second harmonic currents in the field structure windings are found by using the d, q-axes phasor current of frequency 2u as the forcing function. The procedure 1s outlined for the direct axis; it 1s analogous for the quadrature axis. From Eq. (8.9) and (8.10), _4 Yeeneg ~ Beteoneg ~ aeMastaneg * “est eoneg + Men"p nee) (8-56) neg ~ ~*otp-neg ~ at Manta-neg * Mevte-neg * “voty-neg) The voltages on the left-hand side are zero because the damper winding is always shorted, and the de voltage source supplying the field winding is seen as a short-circuit by second harmonic currents. With zero voltages, and knowing that all currents are second harmonics, Eq. (8-56) can be rewritten as two phasor equations gPSulege $2uMen tnee |]. - [Mae] Oop dq-neg (F Stree Ret icene || lpenes $2uMay which are solved for the two phasors I y, Their initial values f-neg’ "D-neg* are found on the basis of Eq. (8.55) as Semeg( ~ “tlt enegh a (8.58) Sp-neg() * “I™{Tp-negh The value L¢_,,,(0) 18 then added to t¢ (0) from Eq- (8.47) for the total initial field current, whereas i ,.,(0) 18 already the value of the total damper current. The phasor currents I, agg Md Tq neg f0F the quadrature axis are obtained analogous to Eq. (8-58), by replacing subscripts f and D with g and Q. Their initial values on the basis of Eq. (8.55) are then 8-32 Agnes’) ~ RelTy neg! ‘neg! ‘q-neg! which are the total initial values since the respective positive sequence eee (8-59) 0) = Re(Y, values are zero. The negative sequence phenomena produce torques which influence the initialization of the mechanical part. Recall that the electromagnetic torque on the electrical side 1s Ayi,-Agty- With both fluxes and currents consisting of positive and negative sequence parts, the total torque can be expressed as the sum of three terms, 1 T +7 +7 ‘gen ~ "gen-pos * "gen-neg * *gen-pos/neg The positive sequence torque was already defined in Eq. (8.49), and the (8-60) negative sequence torque is Tr - 1 - . gen-neg ~ *a-neg *q-neg ~ *q-neg ta-neg (8-61) The third term = de noat Tgen-pos/neg 7 Xa-postq-neg™ Neeneg*q-pos ‘q~pos A (8-62) i-neg” Sq-neg “pos ts an oscillating torque produced by the interaction between positive and negative sequence quantities, with an average value of zero. That it is purely oscillatory can easily be seen since all positive sequence values in Eq. (8.62) are constant, and all negative sequence values oscillate at a frequency of 20. This and other oscillating terms are ignored in the initialization of the mechanical part, where torsional vibrations are not taken into account. The negative sequence torque of Eq. (8.61) consists of a constant part, which must be included in the initialization of the mechanical part, and of an oscillating part with frequency 4o which is ignored. To find the constant part, the fluxes are first calculated as phasors, Saeneg * ba Taq-neg * Mar Te-neg * “av '>neg Ce) - L, + L, + 1 6: Namneg ~ *q Téq-neg * Mag Te-neg * “aq *e-neg ce With the definition of Eq. (8.55) for the instantaneous values of currents 8-33 and fluxes, and after some manipulations of the equations, it can be shown that the constant part is Tyen-neg-constant ~ ®*{Maverage!t™Taq—neg!IMaverage! B(Taq-negt (8-64) with L A - A, +h 8.64b average ” Z “anes + Sq-neg) « 2) The oscillatory part is not needed, but could be calculated from A, mA, - g-neg. ‘d-neg Shut encase oeet] iacory, aja ceca Tpeecees Groce ese | o(oa°2) Identical values for the constant part are obtained from energy balance considerations (uT = power delivered to network + losses in all windings), 1 wo =F {-3|t |? -R) + 2Re + |r. 2R T gen-nog-constant ~ Z {3lTpogl*Bneg Ba) * Teneg!® Be * Tg-neg! 63 + {I 2k + |1, 2R . Pepe bebo Because 3-rd and higher order harmonics are ignored in the armature windings, and 4-th and higher order harmonics are ignored in the windings on the field structure, the initial torque values are not exact. They are good approximations, however, as can be seen from Table 8.1. This table compares the values obtained from the initialization equations with the values obtained from a transient simulation (Fig. 8-9), for the severely unbalanced case described in Fig. 8.6. The constant torque from the initialization procedure is almost identical with the average torque of the transient simulation (difference 1.6%). Fig. 8.9 further shows that the initial torque from Eq. (8-73) can be quite different from the average torque. Table 8.1 also compares the values for the 2-nd and 4th harmonics (not needed in the initialization, though). ‘The values for the 2-nd harmonic agree quite well, Dut not the values for the smaller 4-th harmonic. This is to be expected, because the 4th harmonic torque is influenced by 3-rd harmonic currents in the armature windings, which are ignored in this initialization procedure. ‘The average value in Fig. 8.9 lies not exactly halfway between the maximum and minimum values because the 4-th harmonic 1s phase~shifted with respect to the second harmonic. Tyen -—sinittial value (mm, from Bq. (8.73) / 8-34 average Fig. 8.9 - Torque obtained with transient simulation for case described in Fig. 8.6 Table 8.1 - Electromagnetic torque for test case of Pig. 8.6 (at=46.2963 us) From Fourier Analysis |Relative From Equations Between 80 ms and 100 | Error Torque Component (iim) ms mim) &) pos = 2.953 from (8.49) ‘or (8.50) average neg = 0.017 from (8.64) or (8.66) 3.019 1.62 2nd harmonic 4.673 4.634 0.84 4th harmonic 0.262 0.517 49.32 8-35 The initial zero sequence values are easy to obtain, either from the 4,q,0-transformation of Eq. (8.7), or from the symmetrical component transformation of Eq. (8.38). Physically, both are the same quantities, except that the d,q,0-transformation is normalized and the symmetrical component transformation chosen in Eq. (8.38) is not. Since the d,q,0- quantities are normalized, 100) =F (4,00) + 1,00 + 1,00), (8.678) or 40) = v3 eft...) (8.670) zero. The zero sequence quantities do not produce any torque, and therefore do not influence the initialization of the mechanical part. 8.4.4 Initialization of the Mechanical Part’” The Links between the electrical and the mechanical part are the angle 8,1(0) from Eq. (8-42), which 1s converted to the mechanical side with Eq. (8.25), and the electronagnetic torques T,,, and T,,. from Eq. (8.32). For the generator torque, the constant part is Tgen-constant ~ "gen-pos‘”) + *gen-neg-constant on the electrical side, which is converted with Eq. (8.25) to the mechanical (8-68) side. Since torsional vibrations coming from the oscillating torques of Eq. (8.62) and Eq. (8-65) in unbalanced cases are ignored in the steady-state initializations, the oscillating term is left off in Eq. (8-68). For the exciter torque, the oscillating terms are ignored as well. Then, *) te zero sequence currents in the BPA EMTP can be incorrect (see beginning of Section 8.4). “*) The initial angles in the BPA EMTP can be incorrect in unbalanced cases, because the negative sequence torques are not included in Eq. (8-68) and (8.69). If Table 8.1 4s typical, these torques are very small, however. 8-36 = -v, 1 2 “Texe-constant ~ ~Ye-pos t¢-poa) * 7 Hernegl” Rexe (8-69) with Ty jog from Eq. (8.57)- Without torsional vibrations, the speeds of all turbine-generator masses are one and the same, and the acceleration of each mass is zero. Then Eq. (8-29) simplifies to (xIL0] = (Teucbine! ~ [Tgen/exc! ~ [sere] CED) with w being the synchronous angular speed and [D,.1¢] being the diagonal matrix of self-damping terms D,. The sum of the turbine torques mst, of course, equal the sun of the electromagnetic and speed self-danping torques, so that there 1s zero accelerating torque initially, D Teurpine-t ” Tgen-constant * Texc-constant *¥ 22, (8-71) Eq. (8-71) is used to find the sum of the turbine torques first, and then to apportion the total to the individual stages from the percentage numbers to be supplied in the input (e-g., 30% of torque in high pressure stage, 26% in intermediate pressure stage, etc.). The right-hand side of Eq. (8.70) is then known, as well as the angle of the generator mass from Eq. (8-32a). [K] is singular. Assume the generator to be mass no. k (with 0, known); then the remaining initial angles may be found in 2 ways: (1) Multiply the diagonal element Kj, with an arbitrary large number (e-g. 103°), and reset the k-th right-1 This will, in effect, change the k-th equation to variable 0, = and side value to this number times 0, specified value of 0,- Then solve the systen of linear equations (8.70), preferably with a subroutine for tri-diagonal matrices (required in the time-step loop anyhow). (2) Starting to the left of generator mass k, find the angles of the lower-nunbered masses recursively from tt Rus 1 oy te ou : ‘A-1,4, (8.72a) (RS = right-hand side terms of Eq. (8.70), and starting to the right of generator mass k, find the angles of the higher-numbered masses recursively from 8-37 } Ras 1 em yt + Sek, (8.726) " | ne a 7 ‘4,441 These recursive equations are derived by summing up rows 1,...1, or by summing up rows 1,...n in Eq. (8.70); in either case, most terms on the left-hand side cancel out because of the special structure of [K], as shown after Bq. (8-30). While the oscillating terms of 7... initial angles of the mechanical part, they must be included in initializing and T,,, are ignored in finding the the electromagnetic torques for solving the differential equations in the time-step loop. This is best done using Tyen-totar() ~ Xgl) 140) ~ A460) 4400) (8-73) where the currents are 4400) = 4 (0) + 4. 0) ‘d-pos ‘d-neg. £400) = 15990) + Lg neg? and the initial fluxes are calculated from Eq. (8.10). Similarly, ot (0) 46(0) + Rol QCD}? (8.74) exe-total() * ““¢-pos where 1,(0) = te pos * Le neg 6 8-38 8.5 Transient Solution The numerical methods for the transient solution part are based on [13]. The basic idea is to reduce the machine equations to a three-phase Thevenin equivalent cireult, similar to that of Fig. 8.5 for the steady-state initialization. The equivalent circuit for the transient solution differs from Fig. 8.5 mainly in two aspect: (a) The impedance matrix [Z] of Fig. 8.5 becomes a resistance matrix [R], after integrating the machine equations (8.9) with the trapezoidal rule of integration, and after reducing the seven equations for all windings to three equations for the armature windings. (b) The sinusoidal voltage sources E" of Fig. 8.5 become instantaneous voltage sources which must be updated from step to step. ‘The updating procedure for the voltage sources requires the prediction of certain variables from the known solution at t-At to the yet unknown solution at t. Different prediction methods have been tried over the years, and their behaviour with respect to numerical stability has gradually improved. Some earlier versions of the TYPE-59 synchronous machine model produced too much numerical noise [131], but beginning with version M36, the prediction methods are quite stable and the simulation results are fairly reliable now [132]. Further progress with respect to numerical stability can only be achieved if the overall EMTP algorithm is changed from a direct to an iterative solution in each time step. 8.5.1 Aseune that the solution at t-At 1s already known, and that the solution at t has to be found next. Then the method works roughly as follows: (1) Predict the generator rotor angle f(t) (first predicted variable). (2) Apply the trapezoidal rule of integration to the R-L branches of Pig. 8.2, in the direct axis as well as in the quadrature axis. Conceptually, this converts each R-L branch into an equivalent resistance in parallel with a known current source, as indicated in 8-39 Fig. 8-10(a) and (b). The zero sequence consists of only one branch (Fig. 8-10(c))+ ve (t-bt ty,(t) uj (t-at tu, (t) ea ecstonm dees (a) d-axis (b) qraxis (e) zero sequence Fig. 8.10 - Resistive networks resulting from trapezoidal rule of integration (u = speed voltages defined in Eq. (8.75)) (3) Reduce the d- and q-axis resistive networks of Fig. 8.10 to one equivalent resistance in series with one equivalent voltage source as shown in Fig. 8.11. For this reduction, assume that vg(t) = vg(t-Bt), which is exact 4f the excitation system is not modelled, or use some other prediction (e.g. linear extrapolation). 8-40 5 aS a & e (a) Direct axis (b) Quadrature axis (c) Zero sequence a) Fig. 8.11 - Resistive networks Unfortunately, the speed voltages ug say (8.75) ay Toy at time t are also unknown, but eince fluxes can never change abruptly, their values can be predicted reasonably well. With Predicted values for u(t), ug(t) and ve(t) (2nd, 3rd and 4th predicted variable), the reduction is straightforward. Conceptually, branches M, f, D for the d-axis in Fig. 8.10(a) are paralleled, and then connected in series with the c-branch (analogous for the q-axis). Convert the 3 resistive Thevenin equivalent circuits for d, 4, O-quantities to phase quantities. If this were done directly, then the resulting 3 x 3 resistance matrix would be time-dependent as well as unsymmetric. To obtain a constant, symmetric matrix, the equivalent resistances of the d- and q-axis are averaged, as RR, indicated in Fig. 8.12, and the “saliency terms” 474+ 44(t) and RR, gia, age 8 and e, into = + 1,(© are combined with the voltage sources e, and e, int 8-41. becomes one source (a) direct axis (b) quadrature axis (ce) zero sequence (4) phase quantities Fig. 8.12 - Modified resistive networks one voltage source. This can only be done at the expense of having to use a predicted value for i4(t) and 1,(t) (Sth and 6th predicted variable). After conversion to phase quantities, the d, 4, O-networks become one three-phase network, with three source voltages behind a symmetric, constant resistance matrix [R,oysy]+ (5) Solve the complete network, with the machine representation of Fig~ 8.12(4). In the EMTP, current sources in parallel with [R.o,,] are used in place of voltage sources in series with [R..,;,]+ ‘equiv: (6) From the complete network solution in phase quantities, extract the generator voltages and convert them to d, q, O-quantities. 8-42 Calculate the armature currents in 4, q, O-quantities and the field structure currents, and use them to find the electromagnetic torques of the generator and exciter from Eq. (8-32) at time t. Then solve the equations of the mechanical part. (7) Compare the predicted values of p, ug Ug, 1g) {4 with the corrected values froa the solution of step (6), and repeat steps (6) and (7) 4f the difference is larger than the acceptable tolerance. When returning to step (6), it is assumed that the terminal voltages in phase quantities remain the same. (8) Return to step (1) to find the solution at the next time step. Some of the implementation details, which have been omitted from this brief outline, are discussed next. Variations of the iteration and prediction methods are described in Section 8.5.4. 8.5.2 Tr Consider the equations for the direct axis first, which are obtained from rows 1, 4, 6 of Eq. (8.9) and from Eq. (8.10a) as Val a a ‘a ‘at a rT -- ~ Mee Mee Men [ate/ae] + 10 (8-76a) ° ca May Mep np} {4tp/at} {0 with ug being the speed voltage from Eq. (8-75), or in short-hand notation, ai. v) = - s 8.71 Iv} (R] (4) - (4) (gel + [ed (8.76b) Because of numerical noise problems in pre-M32 versions of the BPA ENTP, this * equation 4s integrated with the “damped trapezoidal rule” of Section 2.2.2"), with a damping resistance matrix [R,] in parallel with [L], ? UBC's PC Version MicroTran uses the normal trapezoidal rule. By setting a= 1 inthe input, the BPA EMTP would use the normal trapezoidal rule as well. 8-43, 1+a,2 @) = pegs ae Gl (8.77) where a is the reciprocal of the damping factor defined in Eq. (2-21). For a= 1 there is no damping, while a = 0 is the critically damped case. In the present version of the BPA EMTP, a default value of (1 + a)/(1- a) = 100 ts used. Applying the damped trapezoidal rule of Eq. (2.20) for v = L di/dt to Eq. (8.76), with v replaced by (uJ-[v]-[R}{1], results in [v(e)] = (ace) + (nteece-aey) - (18) + SE wy} 140091 (8.78a) with the know history term UntseCe-at)] = (-0iR] + BE (Ly) } (4Ceat))-alv(e-ae) Hatuce-ae)]. (8.780) Eq. (8.78a) describes a voltage source [u(t)]+[hist(t-At)] behind a resistance matrix Boogp) = (RI + ue nu. (8.79) Subscript “comp” is used because such equivalent resistive networks are called “resistive companion models” in network theory [133]. For interfacing the synchronous machine equations with the network equations, the field structure quantities are eliminated from Eq. (8.78)- Dropping subscript "comp" and using subscripts d, f, D again, the field structure currents can be expressed from the last two rows of Eq. (8.78a) With [Reggp] ftom Bas (8-79) a8 -a (te) ef Rep) Hist ¢(t-At) We Ce aE = - 7 A(t) (8.80) p(t] Rey Byp| ntst,(t-at) 0 Ray £ which, when inserted into the first row of Eq. (8.78a), produces a single equation for the d-axis, vg(t) = eg - yt g(t) (8-81) with the reduced companion resistance 8-44, “1 x, ee Sep “af Raa ~ Bag Rap! > (8.828) eee fe - fe] the voltage source red eg 7 ugit) + hist (e-se), (8-82b) and the reduced history term red R, tet, - ve(t ‘fe * ep £7 Ye hists = histy - [Rye Rap) (8-82c) Rey Bpp| frist, By predicting the speed voltage u4(t)=-u(t)A,(t), and by assuming that vet) = ve(t-At), the simple resistive network of Fig. 8-11(a) 1s obtained, with a voltage source e, behind the companion resistance Ry. If R<<2L/At in all RL branches of Fig. 8.10(a), then te a" aE Therefore, Eq. (8-81) essentially represents the trapezoidal rule solution of R, (8-834) a voltage source behind the subtraasient reactance Xj". In publications based on [13], Ry is called ay). If the dynamic behaviour of the excitation system is to be simulated as well, then using v¢(t) = ve(t-At) implies a one time-step delay in the effect of the excitation system on the machine. Such a delay is usually acceptable, because At is typically much smaller than the effective time constant between the input and output of the excitation systems. Alternatively, some type of prediction could be used for v¢(t). The derivations for the q-axis are obviously analogous to those just described for the d-axis, and lead to the single equation for the q-axis, ve) = eg Ret g(t)s (8-84) with the voltage source 8-45, red g TUGiE) F hist g (trae) « (8.85) Here, only the speed voltage u(t) = w(t)Ag(t) mst be predicted because the voltage v,(t) 1s zero. The qraxis resistive network 1s shown in Fig. B.11(b). Again, 1f R¢<2L/At in all RL branches of Fig. 8.10(b), then 21," ce ‘a ae Therefore, Eq. (8-84) essentially represents the trapezoidal rule solution of (8.83b) a voltage source behind the subtransient reactance X,". In publications based on [13], R, 19 called ayy. The equations for the zero sequence quantities (row 3 in Eq. (8.9) and Eq. (8-10c)) are also integrated with the damped trapezoidal rule, which leads to vo(t) = histg(e-At) - Rolg(t) (8-86a) with the companion resistance ta Ry R, + Ly (8-86b) and the known history tera histy(t-at) = oe Lg - aR,) tg (t-at) - avy(t-at) (8.86c) The zero sequence resistive network of Eq. (8.86a) with e, = histg(t-At) is shown in Fig. 8.11(c). In publications based on [13], Ry 1s called a,,. The reduced generator equations (8.81), (8.84) and (8.86a) can be solved in one of two ways: (1) Find a three-phase Thevenin equivalent circuit (resistive companion model) for the network seen from the generator terminals, and solve it together with the generator equations. (2) Add the reduced generator equations to the network equations, and solve them simultaneously. ‘The first approach was used in [98]. It has the advantage that iterations can easily be implemented for the correction of predicted values. However, generators must be separated by distributed-parameter lines with travel time for reasons of numerical efficiency, so that an independent 8-46 three-phase Thevenin equivalent circuit can be generated for each generator (otherwise, M generators would have to be interfaced with one 3 x M-phase ‘Thevenin equivalent circuit). If there are no such lines in reality, artificial stub-lines with tAt must be used to separate the generators. This can result in incorrect answers. For this reason, the first approach has been abandoned in the EMTP. With the second approach, there is no restriction on the number of generators connected to the network, or even to the same bus. However, it does require the prediction of certain variables, which makes this approach more sensitive to the accumulation of prediction errors. It is the only method retained in the present BPA EMTP, and only this method is discussed here. To solve the generator equations with the network equations, the generator resistive networks of Fig. 8.12 in 4, q, O-quantities must be converted to phase quantities, which produces a time-dependent and unsynmetric 3 x 3 resistance matrix. To accommodate such matrices would have required a conplete restructuring of the basic (non-iterative) solution algorithm of the EMTP. Instead, an average resistance Ryy 7 Ry + RYD/2 (8-87) is used on both axes. This requires “saliency terms” 4 4(R4-R,)/2 on the dmaxte and 1,(R,-R4)/2 on the qvaxis, which are added to the known voltage sources eg, generators with X" = X,", these saliency terms are practically negligible. eq by using predicted values for 13, 1, (Fig. 8-12). For For the IEEE benchmark model [74] with different values of X," = 0.135 p.u. and X," = 0.200 pus, the companion resistances are Ry = 3.5844 p.u. and R, 75-3103 peu. for At = 200 ys. These values are practically identical with 21,°/at = 3.5810 pus and 2L,"/At = 5.3052 peu., a6 mentioned in Eq. (8-83). The voltage drop across the saliency terms (Ry-R,)/2 would be 20% of the voltage drop across (R,#R,)/2 with At = 200 us. With the average resistance of Eq. (8.87), the modified equations in d, 4) O-quantities become 8-47 vyCt ®a-mod ‘av tg Ct) vq Ct) = ea-moa | ~ Ray ye) (8.882) vot) ey RB tp(e) where ding 4 a tlt) (8.88b) RyRy a +4 ace (8.88c) Predicted values 14, 1, are used in the last two equations, and the voltage sources behind resistances are then converted into current sources in parallel with the resistances, L Sa-source ” Ky °a-mod, We) 4, =e (8-890) J-source Ry 'q~mod” eee (8-89c) 10-source ~ Ky “0 Finally, the d, q, O-quantities are converted to phase quantities with a redicted angle B(t), which produces a resistive companion model with current P 2 sources 1 41-source cosp sing 3 = L = | cos¢p-120°) sinp-120°) 7 (8.90) ces vz ° ad 43-source coscpei20°) sin(gH20") and parallel with 8-48 . (8.91a) where Ry = (Ry + 2R,)/3, R= (Ry ~ Ryy)/3- (8.91) Since this model is identical with the resistive companion model of Eq. (3.8) for coupled inductances, generators are interfaced with the network equations as if they were coupled inductances. The matrix [Ry aus, conductance matrix [G] of Eq. (1.8) once and for all outside the time step J-4 enters the nodal loop, while the parallel current sources are updated from step to step. After the complete network solution has been obtained at time t, the generator phase voltages are converted to d, q, O-quantities, vq cosB_ =cos(B-120°) cos (B+120°) vol = B sinB = sin(B-120°) —sin(B+120°) (8.92) i W212 Wz and the armature currents are found from a (Ope = “Vv /R (8.93) tg 7 (gma ¥q) Pav ig * (047V9) /Ro- The field structure currents are recovered from Eq. (8.80) for the d-axis, snd from an analogous equation for the q-exis, Finally, the fluxes \g, % are calculated from Eq. (8,10a) and (8.106), and the electromagnetic torques from Eq. (8.32), which are then used to solve the mechanical equations as described next. 8-49 8.5.3 Transient Solution of the Mechanical Part ‘The mechanical part is described by Eq. (8.29), which can be rewritten as (31 ($2 + 1 fol + KI (01 = (yee) (8.94a) with the speeds of the aysten of masses ww) = (8.946) and the net torque It, (8.94) net! * [eurpine! ~ [gen/exc! ] provides the only direct link with the electrical part, B which had to be predicted to Tre torque [Ton /exe with another indirect link through 9, solve the electrical part. Applying the trapezoidal rule (or central difference quotients) to Eq- (8.944) and (8.94b) ytelds pay LOD LoCeAOD Tg py LODE BET ¢ pxy LODGED 7 (8.95) (Toe yop (EAE) TT and face] + [wCeraer] [0¢e)] ~ (¢enae)] (8.96) Replacing [0(t)] in Eq. (8.95) with the expression from Eq. (8.96) produces Ze cr + oy + SE ved} CoC) = (egy Ce] + tntse(e-aey) (8.974) with the known history term Thtsecenaey] = (Fe 31-10] - AE eK)} (w(e-aey) - 2px fece-aed) + (Ty_(t-ae)] (8.976) Normally, it 1s assumed that the turbine power is constant. In that case, the torque on each mass 1 is calculated by using predicted speeds u,, 8-50 P ‘turbine, i Teurbine,i "wy C If constant turbine torque is assumed, then Eq. (8.98) is skipped. With the turbine torques from Eq. (8.98), and with the electromagnetic torques at time t already calculated in the electrical part, Eq. (8.97a) can be solved directly for the speeds [u(t)]. ‘The matrices tal = 3 ta) + to) + SE ox) and to) = 2 3] - wy - SE 0K are tridiagonal, and remain constant from step to step. They are triangularized once and for all before entering the time step loop, with a Gauss elimination subroutine specifically written for tridiagonal matrices, which saves storage as well as computer time. Inside the time step loop, the information in the triangularized matrix is used to apply the elimination to the right-hand sides, followed by backsubstitution ("repeat solution", as explained in Section TII.1). It is worth noting that the form of Eq. (8.94) is the same as the system of branch equations for coupled R-L~C branches. In that analogy, the matrix {J] is equivalent to a matrix [L] of uncoupled inductances, the matrix [D] to a matrix (R] of coupled resistances, and the matrix [K] to an inverse capacitance matrix [C]-1 of coupled capacitances. [T,,,] would be equivalent to the applied branch voltages {v], and (w] would be equivalent to the branch currents [i]. Instead of a series circuit, one can also use a parallel circuit representation where [T,,.] becomes a current source (see Section 9.4). 8.5.4 Predicti The synchronous machine code in the BNTP has undergone many changes, especially with respect to the prediction and correction schemes. The presently used schemes, as well as variations of it, are summarized here. 8.5.4.1 Prediction of uw and B The speeds of all masses are predicted with linear extrapolation, (w(t) = 2fu(t-at)] - Cu(t-2ae)]. (8.99) Since the speeds change slowly, in comparison with the electrical quantities, this prediction should be accurate enough. Predicted speeds are needed in 8-51 two places, namely for the prediction of speed voltages (see Section 8.5.4.3), and for the calculation of turbine torques from Eq. (8.98). The accuracy of the predicted generator rotor speed w,,, 18 nore 1aportant because there is no speed voltage correction in the present iteration scheme. The accuracy of the turbine rotor speeds prediction is less important, because the torque calculations Eq. (8.98) are corrected in the iteration scheme of Section 8.5.1, if constant turbine power is assumed (default option in UBC EMTP, only option in BPA EMTP). L£ constant turbine torque is assumed, then the turbine speed predictions are not needed at all. Fig. 8.13 shows the speed and the electromagnetic torque of a generator by itself (no turbine connected to it), which runs unloaded at synchronous speed and ts then switched into a resistance load at t = 0 [134] (data in Table 8.2). The generator slows down very quickly in this case. The curves were obtained with the UBC EMIP without iterations (no return from step 7 to 6 in Section 8.5.1), and verified with a th-order Runge-Kutta-Merson method (agreement to within 4 digits). Both the UBC and BPA EMTP had bugs in the speed calculation, which were not noticed before in cases of small speed changes. They were corrected after J. Mechenbier proved their existence by using principles of energy conservation as suggested by H. Boentg and S. Ranade [134]. The angle 8 of the generator rotor must be predicted so that the d, q, O-networks of Fig. 8.12 can be converted to phase quantities for the complete network solution in step 5 of Section 8.5.1. There is no correction for this conversion in the present iteration scheme. The angle f is also needed for converting the voltage solution back from phase to d, q, O-quantities in step 6 of Section 8.5.1; here, corrected values are obtained from the solution of the mechanical part {f steps (6) and (7) are iterated. In the UBC EMTP, the predicted value for § is calculated from the Predicted speed o,.. with the trapezoidal rule of integration (8.96), B(t) =B(erat) +28 fw Cent) +w (ty). (8-100) 7 (gen ‘gen ¥36 and later versions of the BPA EMIP use a predictor formula suggested by Kulicke [135], which 1s based on the assumption that p is a fourth-order polynomial of t, 8-52 3600 speed (r/min) + | 1800 speed at 1 s: 517 r/min ° 0.5 1.0 —+ es) torque (Mm) 1 0.75, maximum: 0.81227 MNm 0.50 0.25 Fig. 8.13 - Speed and electromagnetic torque of an unloaded generator when ewitched into a resistance load. (a) Speed, (b) Electromagnetic torque. (8.101) 8-53 By using three known values of f at t~At, t-2At, t-3At, and two known values of the speed = Bae se 2 3 + 2agt + 3aze” + daye (8-102) at t-At, t-2ht, the coefficients a9, -+- a, can be found from the solution of 5 linear equations. This is a Hermite interpolation formula and leads to the predictor formula [7; p- 184, P6 in Table 5.1] B(t) = - 9 {B(t-at) - B(e-2he)} + B(e-3At) + Gat {geqCtAt) + gent -288)} (8.103) Table 8.2 - Generator test case no. 1 [134] Ratings: 160 MVA (three-phase), 15 KV (1ine-to-line), wye-connected. Reactances: Kyo eT peusy Ky! = 0.245 peus, X4" = 0.185 pu. XK, 7 1:64 pus, X," = 0-185 peu (no grinding) X, = 015 puss Kg = 0-16 peu Time constants: Ty," = 5-9 8, Ty,” = 0-03046 5 Tyo” = 0-075 8 Resistances: R, = 0.001096 p-u- R, = 1089 in steady state (no effect; added because load sone versions cannot handle isolated generator) Rigad ~ 19 after switching at t = 0. Moment of inertia: J = 999.947 (N-m)s?. One pole pair- Terminal voltage: V, = 12.247 2316 (symmetrical in 3 phases). Step size At = 200 us- £ = 60 He KV (peak) in steady state in phase 1 With coefficients a, .-. a known, a predictor formula for w= dp/dt for 0 ih igen use in the speed voltages could be written down with Eq. (8-102) as well, 8-54 bt w(t) = L4Atw(t-at) + 17Atu(e-2at) - 27—(e-Ae) + 24p(e-2ae) + 3p(t-3At) (8-104) The BPA EMTP uses the predicted speed from Eq. (8.99), though. It is not clear whether the 4-th order predictor of Eq. (8-103) is really superior to the predictor of Eq. (8.99). 8.5.4.2. Averaging of d- and q-Axis Companion Resistances Instead of an average resistance (Ry +R,)/2 in Fig. 8.12, the 139 version of the BPA EMTP uses Ry on both d- and q-axes. To compensate for this, a term (RO-R,) + 4, ts added to the voltage source on the q-axis, and no compensating term is needed on the d-axis. Whether this method is better than the averaging prodecure of Fig. 8.12 is unclear. Both procedures are special cases of a class of averaging methods discussed in [136]. 8.5.4.3. Prediction of 14, 1 The armature currents 14, 1, mst be predicted so that the saliency terms 14(Ry-R,)/2 and 1,(R,-Ry)/2 can be coubined with the known voltage sources e,, ¢, (Fig. 8-12). No correction is made for this in the present iteration schene. Note that the saliency ters are practically zero if X," = X,"+ In the UBC version and in BPA versions until M32, the predicted described currents 14, 4, are also used to find predicted speed voltages, in the next section. In the UBC EMTP, linear extrapolation is used, A(t) = 2L(t-at) - 1(t-2at), (8-105) where "i" is either 1, or 1,. The BPA version uses a linear three-point predictor formula which smoothes numerical oscillations. With the current s(e-280) = Tie) (oreasered) sense) s(e-38e) Fig. 8.14 - Linear prediction with smoothing 8-55, known at t-At, t-2At and t-3At as indicated in Fig. 8.14, averaged values axe first found at the two midpoints by linear interpolation 1-2 baw = A(e~3ae) + £(er2Ae) ue a baey = £eertae) + serae) Then a straight line is drawn through the two midpoints, with a slope of A(tat) - £(t-3ae) —m to predict the current 4(t), slope = a(t) = B acerney +f ace-aaey = 2 1¢e-3ae) (8-106) This linear prediction with smoothing is conceptually similar to fitting a straight line through three points in the le: squares sense. Such a straight Line least square fitting would have the sane slope, but a value at t-2bt of {i(t-3At) + i(t-2At) + A(t-At)}/3 instead of {41(t-3at) + 2i(t-2at) + A(t-At)}/4 in Fig. 8.14, which would yield a predictor 4(e) =f acerney +} a¢e-2aey - F a¢e-3ae) (8.107) Which predictor performs best is difficult to say. All predictor formulas discussed in this section depend solely on past points, and not on the form of the differential equations for ig, 1,. Eq. (8.76), and an analog equation for the q-axis, were tried at one time as Euler predictor formulas, but they performed worse that the predictors discussed here. It might be worth exploring other predictor formulas, because the accuracy of the solution depends primarily on the prediction of 14, 1,, espectally if the speed voltages are calculated fron 14, 1, as well, as discussed in the next section. One could use Eq. (8-103), for example, by replacing g with i and w with di/dt calculated from Eq. (8.76)- Fig. 8.15 shows the current in phase 1 after a three-phase short-circuit of a generator with unrealistically low armature resistance R,=0.0001 p.u. The data for this case is summarized in Table 8.3. Since speed changes were ignored, the only predicted values are 14, 1,, a8 well as speed voltages in the BPA EMTP. An improved prediction scheme was recently implemented in MicroTran, which brings the answers in Fig. 8.15 and 8.16 close to those obtained with iteration schemes. 8-56 40 i, (ka) ie of -20 40 +, o 10 20 — t(ms) 4, (ka) ‘4 UBC non-iterated x UBC iterated & MicroTran 9 exact © BPA M39 40 $4, 480 490 500 | —— t(ms) 40} + 970 980 990 —> t(ms) Fig. 8.15 - Current in phase 1 after a three-phase short-circuit. R, = 0.0001 p.u. 8-57 Table 8.3 - Generator test case no. 2 Ratings 400 MVA (three-phase), 18 KV (1ine-to-1ine), wye-connected = 0.92 pus, X4' = 0-18 peu, X, Reactances: g” 7 0-161 poue 0.161 p-u- (no g-winding) a = 0.748 pur, Xy x x X_ = Xp = 0-082 pou. Tr 1, Time constants: += 6.68, T," = 0.05 5 ‘do "= 0.05 s Resistances: 10~" p.u. (unrealistically low value) 1073 p.u. (more realistic value) Bioag * 12 Moment of inertia: J = = (constant speed) Terminal voltage: V, = 4.926 eS 731 RV (peak) in steady state in phase 1 (symmetrical in 3 phases). Step size At = 200 us. f = 60 Ba. Disturbance: Three-phase short-circuit at terminals at t In such a case with low damping, the errors caused by the prediction do accunmilate noticeably if the simulation runs over 5000 steps to t,,, = 1s- The errors are decreased if the complete network solution is iterated (not yet available as an option in the production codes of the EMP). For comparison purposes, the exact solution 1s shown as well, which was found for ig tg, with the etgenvalue/eigenvector method discussed in Appendix T-1, and then transformed to phase quantities with @ from Eq. (8-23). Fig. 8-16 shows the results if the armature resistance 1s changed to a more realistic value of R, = 0.001 p.u. As can be seen, the answers are now closer to the exact solution. 8.5.4.4. Prediction of Speed Voltages Starting with M32 of the BPA ENTP, the speed voltages uy, u, of Eq- (8.75) are predicted in the sane way as 14, 1, with Eq. (8.106). In some of these versions, the prediction was done in a synchronously rotating reference frame, and then converted directly to phase quantities without going through 4, (ka) 4(ka) 4, (ka) -40 40 40 20 -40 Fig. 8-58 fee cee) 4 UBC non-iterated x UBC iterated & MicroTran © exact 480 490 500 — t(ms) & y 970 980 990 — t(ns) 8.16 ~ Current in phase 1 after a three-phase short-circuit. R,= 0.001 p.u. 8-59 4, q-axes parameters. This has been abandoned in Feb. 1986, and the speed voltages are now again predicted in d, q-quantities because the latter turned out to be superior when applied to test case no 1 of Table 8.2. In pre-M32 versions of the BPA EMTP, and in the (unreleased) UBC version with synchronous machines, the speed voltages u. are not ‘a ‘a predicted explicitly. Instead, the predicted currents 14(t), 14(t) and the or way and us = ai, Predicted speed w,,,(t) are used to calculate the speed voltages from Eq. (8-10a) and (8.10b). The field structure currents appearing in these equations are expressed as a function of i, with Eq- (8-80), which leads to 4 the expression reduced agit) = Ly 14(0) + gig (8-108) with the known reduced inductance 1 R, reduced ee Be) ‘at Ly T Ugo yg Bal | 7 (8-109) 20 pp| a0 and the known flux 249 for zero current (14 = 0), R, dst y—v, (t) re ep] £ Ve Nao Mas Man) |, og (8.110) 20 ®pp| — |nist, D reduced The reduced inductance Ly is practically identical with L," if RC<2L/At. For the IEEE benchmark model {74} with At=200 us, reduced ‘a based on [13], ul, ol = 0.135129 p.u. compared to ul,” = 0-135 p.u. In publications reduced reduced a ie called ~ay, and ul, is called ay). 8.5.4.5. Iteration Schemes Up to now, the complete network solution is direct, without iterations. The iteration scheme of Section 8.5.1 does not repeat the network solution, and predicted values are therefore never completely corrected. There is only one exception, namely the three-phase short-circuit at the generator terminals with zero fault resistance. In that case the terminal voltages are 8-60 always zero, and going back to step 6 in the iteration scheme of Section 8.5.1 should be a complete correction of all predicted values. It is doubtful whether the predictors can be improved much more. Further {mprovements can probably only be made if the network solution is included in the iterations as well, This could be a worthwhile option, not only for machines, but for other nonlinear or time-varying elements as well. 8.6 Saturation Saturation effects in synchronous machines can have an influence on load flow, on steady-state and transient stability, and on electromagnetic transients. While transformer saturation usually causes more problems than machine saturation (e-g., in the creation of so-called “temporary overvoltages » there are situations where saturation in machines mst be taken into account, too. To model machine saturation rigorously is very difficult. Tt would require magnetic field calculations, e.g. by finite element methods [181], which is already time-consuming for one particular operating condition, and practically impossible for conditions which change from step to step. Also, the detailed data for field calculations would not be available to most EMTP users. An approximate treatment of saturation effects is, therefore, commonly accepted. The modelling of saturation effects is discussed in four parts, (a) basic assumptions, (b) saturation effects in steady-state operation, and (c) saturation effects under transient conditions, and (4) implementation in the ENTP 8.6.1 Bi The data which is normally available is the “open-circuit saturation curve" (Fig. 8-17), which shows the terminal voltage as a function of the field current for open-circuited armature windings (no-load condition). In the transient simulation, a flux-current relationship is required, rather 8-61 than V = f(tg). This is easily obtained from Eq. (8.9), which becomes Vg 7 Why (8-111a) (can be re-labelled as flux) — feno load Fig. 8.17 - Open-cireuit saturation curve for balanced, open-circuit steady-state conditions, where both 4, and the transformer voltages dh,/dt, ad,/at are zero. Since v, ts equal to /3V, puss this equation can be rewritten 1. wn, (8-111) a-rus ~ “a where V,_gys 18 the RMS terminal voltage of armature winding 1 (1ime-to-ground RMS voltage for wye-connected machines). It is therefore very simple to re-label the vertical exis in Fig. 8.17 from voltage to flux values with Eq. (8-111). The ‘uration effects in synchronous machines do not produce harmonics during balanced steady-state operation, because the open-circuit saturation curve describes a de relationship between the de flux of the rotating magnets (poles) and the de field current required to produce it. The magnitude of the de flux determines the magnitude of the induced voltages in the armature, while the shape of the flux distribution across the pole face determines the waveshape of the voltage. If the distribution is sinusoidal, as assumed in the ideal machine implemented in the EMIP, then the voltage will be 8-62 sinusoidal as well. In reality, the distribution is distorted with “space harmonics", and it is this effect which produces the harmonics in synchronous machines. There are many different ways of representing saturation effects [182], and it is not completely clear at this time which one comes closest to field test results. More research on this topic is needed. At this time, the representation of saturation effects in the EMIP is based upon the following simplifying assumptions: a The flux Linkage of each winding in the d- or q~ axis can be represented as the sum of a leakage flux (which passes only through that winding) and of a mtual flux (which passes through all other windings on that axis as well), as illustrated in Fig. 8.18, rea tal (8-112) where A, = leakage flux unaffected by saturation, dg 7 mutual flux subjected to saturation effects. In reality, the leakage fluxes are subjected to saturation effects as well because they pass partly through iron [180], but to a mich lesser degree than the mitual flux. Saturation effects are therefore ignored in the leakage fluxes. The data is not available anyhow if only one saturation curve (open-circuit saturation curve) is given. In terms of equivalent circuits, this assumption means that only some of the inductances are nonlinear (shunt branch in star point in Fig. 8.2), while the others remain constant. The degree of saturation is a function of the total air-gap flux Linkage op hear eChe) (8-113a) with z z 2. (8.113) and Agaru 7 Maula + ty +4595 Magu = Myultg # fy + 4g)» (B-113e) 8-63 mutual flux A, Fig. 8-18 - Leakage fluxes and mutual flux where eubscript "m” indicates mtual, and “u” indicates unsaturated values. In Eq. (8-113c) it is important to use the proper mutual inductances for the representation of the mitual flux. This leads back to the data conversion problem discussed in Section 8-2. If Canay's characteristic reactance X, is not known, then assume k = 1 in Eq. (8-14b), and use 5 MAL - by, qu" “a from Eq. (8-20a) and (8.20c). In this cs % , the equivalent star circuit of Fig. 8.2 shows the correct separation into the mtual inductance turation) and into the M, = Mgy OF May for the mitual flux (subject to Leakage inductances for the leakage fluxes (linear d-, f-, and D-branches). If Canay's characteristic reactance is used, then Fig. 8.2 can no longer be used, as explained in Section 8.6.5. Only one flux, namely the total air-gap flux, is subjected to saturation. The saturated mtual fluxes hyy+ Nyy OR both exes are found from thefr 8-64 unsaturated values by reducing them with the same ratio (similar triangles in Fig. 8.19), A mq ‘mq-u Fig. 8.19 - Unsaturated and saturated mtual fluxes a, a, eg" ney “Gt dD, oot 5 (8-114) deo heer dees hes This assumption is based on the observation that there is only one mutual flux, which Lines up with the pole axis {f 4, is very enall, and which will shift 0 one side of the pole as A,, increases (Fig. 8.20). (a) flux from t¢ alone (g,"0) (b) flux from i¢ and stator currents Ag #0 Fig. 8.20 - Flux in turbogenerator [181]. © 1981 IEEE 8-65, 4. Saturation does not destroy the sinusoidal distribution of the magnetic field over the face of the pole, and all inductances therefore maintain their sinusoidal dependence on rotor position according to Eq. (8-5)+ 5. Hysteresis is ignored, while eddy currents are approximately modelled by the g-winding, and maybe to some extent with the D- and Q-windings. More windings could be added, in principal, to represent eddy currents more accurately. 8.6.2 Saturation in Steady-State Operation At this time, the saturation effects are only modelled correctly in the ac steady-state initialization if the terminal voltages and currents are balanced. More research is needed before saturation can be represented properly in unbalanced cases. As explained in Section 8.4, the initialization of the machine variables follows after the phasor steady-state solution of the complete network. The initialization for balanced (positive sequence) conditions is described in detail in Section 8.4.1, and only the modifications required to include saturation effects will be discussed here. ‘The nonlinear characteristic of Fig. 8.17 makes it impossible to use the initialization procedure of Section 8.4.1 in a straightforward way. To get around this problem, it is customary to use an “equivalent linear machine” in steady-state analysis which gives correct answers at the particular operating point and approximate answers in the neighbourhood. This equivalent linear machine is represented by a straight line through the operating point and the origin (dotted line in Fig. 8.21). Whenever the operating point moves, a new straight line through the new operating point mst be used. The concept of the equivalent linear machine is used in the EMTP as follows. 1. Obtain the ac steady-state solution of the complete network. From the terminal voltages and currents of the machine (positive sequence values), find the internal machine variables with the method of Section 8.4.1. 8-66 i. air gap equivalent 4] tine Tinear machine l A Fig. 8.21 - Linearization for steady-state analysis Assume that the machine operates in the unsaturated region at this time. Determine the total magnetizing current ee 24 (Cae 2 7 (gt te +A)? + (WEE) Ly tty tt)? > (8-115a) being zero for balanced condition: Boh Eq. (8.115) assumes 11 (same for quadrature axis). If any wth tp, 4, 4, with tp, 1, 49 turns ratios of N, other turns ratios are used, the first term would be Mautaa 7 Mauta + Mag-ute + Map-utp» (8-115) and the second term i, (8-115) Me el Meee Meet Med quemg ~ “quiq” “qg-we qtu with ei 2 2 Masten 2 1 Maytag? + Sgateg) (8-154) Find the operating point on the nonlinear characteristic of Fig. 8.21. Ig this point lies in the linear region, then the initialization is complete. Otherwise: 8-67 calculate the ratio K from Fig. 8.21, K = AB/AC (8.116) and multiply the unsaturated mutual inductances with this ratio to obtain the saturated values of the equivalent linear machine, Mg = KMay 7 Mg = K'Mgy (8.116) Use these values to repeat the initialization procedure of Section 8.4.1. Then re-calculate the magnetizing current from Eq. (8.115). If it agrees with the previously calculated value within a prescribed tolerance, then the initialization is finished. If not, repeat step 3. Convergence is usually achieved with 1 to 2 iterations of step 3. This fast convergence is not surprising. P. Barret, B. Blengino and D. Souque of Electricité de France have shown that the saturation correction is actually non-iterative if the total flux is corrected for the saturation. It does become iterative, however, if the fluxes on the direct and quadrature axes are corrected separately. With Va =~ Raig = Og Vq = 7 Rag + Og ‘a from Eq. (8.9), and with separation of the fluxes into leakage flux and mutual air gap flux according to Eq. (8.112), lg = lpg + OLyig Ohg = Ogg + Obpig we get the mutual air gap fluxes directly from the terminal voltages and currents: hag = Vq + alg - OLghg , hag = "Va ~ Raia - Oleg « Both mutual air gap fluxes do not depend on saturation. How the total flux splits up into the direct and quadrature axes components does depend on saturation, however, because the quadrature axis position depends on saturation. ‘If we apply the saturation correction to the total flux, then it does not matter where it is located with respect to the quadrature axis, as long as the magnitude is correct. In that case we can replace the quadrature and direct axes quantities with the real and imaginary quantities 8-674 referred to the reference frame used for the network solution. Then we can rewrite the above equation as he reat t Ihe teasinnr = ys (954.5%) 1 F - or ky = [V+ (Ry 45%) I]. This value is used in Fig. 8.21 (rescaled from flux 4, to induced voltages iy) to get the ratio K directly from Eq. (8.116a), without any need for iterations. In the BPA EMTP steady-state solution, machines are now represented as voltage sources at the terminals, and the terminal currents are obtained from that solution. With terminal voltages and currents thus known, their positive sequence components can be calculated and then used to correct the internal variables for saturation effects. Since this correction does not change the terminal voltages and currents, the complete network solution does not have to be repeated in step 3. This will also be true in future versions of the EMIP, where the machine will be represented as symmetrical voltage sources behind an impedance matrix. Again, the terminal voltages and currents and their positive sequence components will be known from the steady-state solution. In unbalanced cases, the present representation will give incorrect negative sequence values, while the future representation will produce correct values. How to use these negative sequence values in the saturation corrections has not yet been worked out. Since they produce second harmonics in the direct and quadrature axes fluxes, it may well be best to ignore saturation effects in the negative sequence initialization procedure of Section 8.4.2 altogether. The equivalent linear machine produces correct initial conditions for the different model used in the transient simulation, as can easily be verified if a steady-state solution is followed by a transient simulation without any disturbance. 8-68 The equivalent linear machine described in Section 8.6.2 cannot be used in the transient solution, because the proper linearization for small disturbances (as they occur from step to step) is not the straight line 0-C in Fig. 8.21 (“linear inductance"), but the tangent to the nonlinear curve in point C (“incremental inductance”). The saturation effect enters the transient solution discussed in Section 8.5 in two places, namely through the speed voltages and through the transformer voltages. Consider the direct axis equations (8.76) first, which can be rewritten as a| ‘4 a [om “hag aa t= 1 - se] de | ae] tea ft] O° +| 0 (8-117) dep Mad ° for the d,f,D-quantities if each flux linkage is separated into its leakage flux and the common mutual flux, dat at nw? de 7 Age * nae (8.118) Ap 7 yy * Ana? assuming turns ratios of Ny:NeiN, = 1:1:1 (analogous for quadrature axis). Only the terms with Ay and Ay in Eq. (8-117) are influenced by saturation, and only those terms are therefore discussed. Consider first the speed voltage term -wij, in Eq. (8-117), which ts properly corrected for saturation by simply using the correct saturated value Ngq’ The saturation correction has already been described in fq. (8-114), and is conceptually the same as the one used in Eq. (8-116) for the steady-state solution. Since the transient solution works with predicted values of speed voltages wry and uh,, as explained in Section 8.5.4.4, they are used directly in Eq. 8-117 (not split up into two terms). Next consider the transformer voltages -[4\.4/4t] in Eq. (8.117), where incremental changes (“incremental inductances”) are important. By using 8-69 the tangent of the nonlinear characteristic in the last solution point, one can linearize the flux-current relationship to a” Mnee * Metopeta a dm slope= x 2 pes Le OPENS ope i lee solution at t-At Fig. 8.22 - Linearization around last solution point with M, being an incremental inductance (Fig. 8-22). This equation can be used over the next tine step, because the fluxes change only very slovly with typical step sizes of 50 to 500 ys. In the EMTP implementation, the problem is even simpler because the saturation curve is represented as a two-slope piecewise linear curve. In that case, the linearization of Eq. (8.119) changes only at the instant where the machine goes into saturation, and at the instant when it comes out again. With Eq. (8.113) and (8.115), the unsaturated total flux is Meu * Maule (8-120) which, inserted into Eq. (8-119), produces Aa 7 Mxnee * Panu (8-121a) with the ratio between incremental inductance M,,,, and linear (unsaturated) pe inductance My,» 8-70 M, p = ‘slope au After saturation has been defined for the total flux, it must be (8.121b) separated into d- and q-components again. With assumption (3) from Section 8.6.1, and with Fig. 8.19, Yad ~ Menee-d + P*na-u? *oq “ Senee-q * PAnq-u , (8.122a) where Xenee-d ~ Meneo¥ * Mneo-q ~ MeneeSt*¥ } ¥ If Eq. (8-117), and the analogous equation for the quadrature axis, are snd-y) (8-122) solved with the trapezoidal rule of integration, then the transformer voltage term affected by saturation, (Wag) = “al Aggl/at is transformed with Eq. (8-122) into 2b Uyglt)] = - Fe (Dugg (t)] - Dygay (tae) } 2 ~ ae [nee alt] — Oyneena(t-Atd]} — (vgg (tae) ] (8.123) This equation shows how the transformer voltages must be corrected for saturation effects: (a) mitiply all mutual inductances by the factor b, and (b) add correction terms to account for the variation of the knee fluxes Menee-al? 284 Mnee-q! 8.6.4 Impl Saturation effects were modelled for the first time in the M27 version of the BPA EMTP, based on the concept of two independent saturation effects, one in the direct axis and the other in the quadrature axis. This was replaced with a newer model in the M32 version, which was essentially the model discussed here. It was not quite correct, however, because the 8-71 correction terms in Eq. (8.123) related to the knee fluxes were not included. The model described here was implemented for the first time in the DCG/EPRI version to be released in 1986. The open-circuit saturation curve is approximated as a two-slope piecewise linear characteristic (0-1 and 1-2 in Fig. 8.22). The number of linear segments could easily be increased, but a two-slope representation is usually adequate. 8.6.4.1 Steady-State Initialization The initialization procedure is only correct for balanced networks at this time. The extension to unbalanced cases is planned for the future. Until this is done, some transients caused by incorrect initialization can be expected in unbalanced cases. Hopefully, they will settle down within the first few cycles. The initialization follows the procedure of Section 8.6.2. For the Feactances X,,X,, which consist of a constant leakage part and a saturable mutual part, x, TX tus RR to, (8.124) unsaturated values My, Mj, are first used to obtain the internal machine variables with the method of Section 8.4.1. If the resulting magnetizing current lies in the saturated region, then the mutual reactances M,, Mj in Eq. (8.124) must be corrected with Eq. (8-116). The calculation of the internal machine variables is then repeated with saturated reactances one or more times, until the changes in the magnetizing current become negligibly small. With the two-slope piecewise linear representation implemented in the EMTP, the ratio K needed in Bq. (8.116) is c= Mstopete * Mence rey ate with the meaning of the parameters show tn Fig. 8.22, and with t, calculated from Eq. (8-115). 8-72 8.6.4.2 Transient Solution Saturation effects in the time step loop are modelled according to Section 8.6.3. The coefficient b of Eq. (8-121b) is set to 1.0 in the unsaturated region, and to M,1,,6/ solution moves from one region into the other, tt is reset accordingly. Mj in the saturated region. Whenever the This coefficient b affecte the values in the equivalent resistance matrix [R,gzy] Of Fa- (8.918) and in the history term matrix of Eq. (8.82c)- To include this coefficient, the inductance matrix of Eq. (8.76) is split up into md-u “ad-u "nd-u ‘24 => | ay Cage Eada (8-126) af L, Tada Tada bna-w ‘| (analogous for quadrature axis). Whenever b changes , [L] is recalculated and then used to recalculate (Rjgyyy] and the history term matrix of 8. (8.82c). With the two-slope representation implemented in the EMP, there are only two values of b, and the matrices could therefore be precalevlated ‘slope! Mau" The major effort lies in the re-triangularization of the network conductance matrix [G] of Eq. (1.8), however, which contains [R, ‘equiv changes whenever the machine moves into the saturated region, or out of it. outside the time step loop for the tvo values of b = 1 and b = ¥, J" and therefore An additional modification is required in the calculation of the history terms with Eq. (8.78). As shown in Eq. (8-123), the knee fluxes A, 4(t) and Aynee(t~At) must now be included. Since the trapezoidal rule of Antegration is not very good for the calculation of derivatives, the knee fluxes are included with the backward Euler method. First, the knee fluxes Ayneed(t) 894 Ayneeg(t) are predicted, using the three-point predictor of Eq. (8-106). Then the trapezoidal rule expression pale 1Ce) Eas (ees) is replaced with the backward Euler expression 8-73 “ predicted = {0 cea (E)] - DneenaGt O00] + (8-127) and the voltage term [v,4(t-At)] 1s replaced by a voltage term which excludes the knee flux. 8.6.5 S If saturation 1s ignored, then it does not matter whether Canay's characteristic reactance is used or not, because it only affects the data conversion part. With saturation included, however, the nonlinear inductance can only be identified as the shunt branch M, in Fig. 8.2 if k = 1 in Eq. (8-14b). If Canay's characteristic reactance is known, then k #1. This factor k must then be removed again from the rotor quantities in Bq. (8-15a), by multiplying the second and third row and column with its rectprocal value, Xa ‘d cae oy a = - Clee, 62M = (8-128) = vw ce vz = mM Mt 2 = i oo» ey ype | | eto where Bo (o.129) and where M,, Lee, and Lpp, are the modified parameters straight out of the data conversion routine of Appendix VI.4. As explained in the text between Eq. (8-17) and (8-18), the factor /3//Z in Eq. (8-128) 1s needed to produce can the warns ration of Ny:tpiy = Iels1. Only with turns xottoe of fovea be sepacaved into their aain and leakage parts. Te clzouit of Fig. 9.29, which 1a equivalent to B. (6128), hae the correct separation {nto the mutant inductance of, <2 Ma, subjected to saturation (Eot the autual #108), 8-74 Fig. 8.23 - Equivalent circuit for direct axis with identity of leakage and main fluxes restored from Fig. 8.2 and into the linear leakage inductances in the three branches d, f, D. For the quadrature axis, Fig. 8.2 can still be used, with M, being the nonlinear inductance, because Canay's characteristic reactance cannot be measured on that axis (current split between g- and Q- windings unmeasurable because both windings are hypothetical windings). Most EMTP users will not know Canay's characteristic reactance because it is not supplied with the standard test data. Therefore, it has not yet been included in the saturation model in the EMTP, e.g. in the form of Fig. 8.23, because of lower priority compared to other issues. When it is implemented, one would have to decide whether the inductance c2M, - cM, which is mutual to both f- and D- windings, should be constant or saturable as well. 9. UNIVERSAL MACHINE Co-author: H.K, Lauw The universal machine was added to the EMTP by H.K. Lauw and W.S. Meyer [137,140], to be able to study various types of electric machines with the same model. It can be used to represent 12 major types of electric machines: (2) synchronous machine, three-phase armature; (2) synchronous machine, two-phase armature; (3) induction machine, three-phase armature with squirrel-cage rotor; (4) induction machine, three-phase armature and three-phase wound rotor; (5) induction machine, two-phase armature; (6) single-phase ac machine (synchronous or induction), one-phase excitation; (7) same as (6), except two-phase excitation; (8) dc machine, separately excited; (9) de machine, series compound (long shunt) field (10) de machine, series field; (11) de machine, shunt compound (short shunt) field; (12) de machine, shunt field. The user can choose between two interfacing methods for the solution of the machine equations with the rest of the network. One is based on compensation, where the rest of the network seen from the machine terminals is represented by a Thevenin equivalent circuit, and the other is a voltage source behind an equivalent impedance representation, similar to that of Section 8.5, which requires prediction of certain variables. The mechanical part of the universal machine is modelled quite differently from that of the synchronous machine of Section 8. Instead of a built-in model of the mass-shaft system, the user must model the mechanical Part as an equivalent electric network with lumped R,L.C, which is then solved as if it were part of the complete electric network. The electromagnetic torque of the universal machine appears as a current source in this equivalent network. 9.1 Basic Equations for Electrical Part Any electric machine has essentially two types of windings, one being stationary on the stator, the other rotating on the rotor. Which type is stationary and which is rotating is irrelevant in the equations, because it is only the relative motion between the two types which counts. The two types are: (a) Armature windings (windings on "power side” in BPA Rule Book). In induction and (normally) in synchronous machines, the armature windings are on the stator. In de machines, they are on the rotor, where the commutator provides the rectification from ac to de- (b) Windings on the field structure (“excitation side” in BPA Rule Book). In synchronous machines the field structure windings are normally on the rotor, while in de machines they are on the stator. In induction machines they are on the rotor, either in the fora of a short-circuited squirrel-cage rotor, or in the form of a wound rotor with slip-ring connections to the outside. The proper term is “rotor winding” in this case, and the term “field structure winding” is only used here to keep the notation uniform for all types of machines. These two types of windings are essentially the same as those of the synchronous machine described in Section 8-1. Tt is therefore not surprising that the system of equations (8.9) and (8.10) describe the behaviour of the universal machine along the direct and quadrature axes as well. The universal machine is allowed to have up to 3 armature windings, which are converted to hypothetical windings 4,q,0a ("a" for armature) in the sane way as in Section 8.1. The special case of single-phase windings is discussed in Section 9.3. The field structure ts alloged to have any number of windings DL, D2, --. Dm on the direct axis, and any number of windings Ql, 02, ... Qn on the quadrature axis, which can be connected to external circuits defined by the user. In contrast to Section 8, the field structure may also have a single zero sequence winding Of ("f" for field structure), to allow the conversion of three-phase windings on the field structure (as in wound-rotor induction machines) into hypothetical D,Q,0-windings- 9-3 With these minor differences to the synchronous machine of Section 8 in mind, the voltage equations for the armature windings in d,q-quantities become | with w being the angular speed of the rotor referred to the electrical side, and in zero seq oa The voltage equ: fro] 2 ars 2. and Fun, oo - +f Fo ‘al [tq Ag [tery fea Ry tog ne dig,/dt- ations for the field structure windings are DL i i] pal i toa] om oa F QL Qu ‘Qh Ryo Sqat doe =~ Bostog = at (91a) (9-1b) (9-28) (9-2b) (9-20) ‘The flux-current relationships on the two axes provide the coupling between the armature and field structure sides, (9-3a) ty Mogi Mago eee Moon 4g ra Moan Tar Mange *** Mann |} a1 oa - Haga Moig2 Too tee Moan ign 7 (9.3b) Maan Maron Mazon *** Yan J fon with both inductance matrices being symmetric. The zero sequence fluxes on the armature and field structure side are not coupled at all, oa ” Loatoa (9.3c) Doe * Lostoe (9.3¢) The universal machine has been implemented under the assumption that the self and mututal inductances in Eq. (9.3a) and (9.3b) can be represented by a star circuit if the field structure quantities are referred to the armature side, as shown in Fig. 9.1. This assumption implies that there is only one mutual (or main) flux which links all windings on one axis (A, in Fig. 8.18), and that the leakage flux of any one winding is only linked with Fig. 9.1 - Star circuit representation of coupled windings in direct axis (analogous in quadrature axis). that winding itself, Strictly speaking, this is not always true. For example, part of the leakage flux of the field winding with respect to the armature (Age in Fig. 8.18) could be linked with the damper winding as well, but not Linked with the armature winding, which leads to the modified star circuit of Fig. 8.23 (synchronous machines) or Fig. 9.2 (induction machines). The data for such models with unequal mutual inductances is seldom available, however (e.g. Fig. 8.23 requires Canay's characteristic reactance, which is not available from standard test data). ‘The star circuit is therefore a reasonable assumption in practice. At any rate, the code could easily be changed to work with the self and mutual inductances of Eq. (9.3) instead of the star circuit of Fig. 9.1. With the star circuit representation of Fig. 9.1, the flux-current equations (9.3a) can be simplified to da = Leata + Ana? p. * “oriba * na’ (9.4a) Lipnttm * Sma with (9.4b) Need eget ee where the prime indicates that field structure quantities have been referred to the armature side with the proper turns ratios between d and Dl, d and D2, +++ and D,. All referred mutual inductances are equal to My in this representation, and the referred self inductances of Eq. (9.3a) are related to the leakage inductances of the star branches by Lh = Lip +, bi ~ Yer * Ma (9.5) Tm “pm * Ma ‘The voltage equations (9.1) and (9.2) are valid for referred quantities as well, if Rygripg, ++. are replaced by R},,i3), ... The quadrature axis equations are obtained by replacing subscripts d, D in the direct axis equations with 4,Q. 9-6 In the BPA EMTP Rule Book, the turns ratios are called “reduction factors", and the process of referring quantities to the armature side is called “reduction” (referring quantities from one side to another is discussed in Appendix IV.3). 9.2 Determination of Electrical Parameters By limiting the universal machine representation to the star circuit of Fig. 9.1, the input parameters are simply the resistances and leakage inductances of the star branches and the mtual inductance, e-g-, for the direct axis, Rar bag Bo Bir Ber Spe My (analogous for the quadrature axis), and for the zero sequence on the armature and field structure side, “oa? Roer Lgee If the armature leakage inductance L, is known instead of the mutual inductance, then find M from Eq. (9.5), eee Me eT Te Cr ne a a If neither Ly nor M is known, then use a reasonable estimate. The BPA EMIP Rule Book recommends Lag TOL Uys byg = Oe Lys (9-6) which seems to be reasonable for round rotor synchronous machines, while for salient pole machines the factor is closer to 0.2 than to 0.1. Compared to the large value of the magnetizing inductance of transformers, the value of 9-7 the mtual (or magnetizing) inductance Mj, M, from Eq. (9.6) (90% of self inductance) is relatively low because of the air gap in the flux path. Compared to the (w+1)(s+2)/2 inductance values in Eq. (9.3a), the star circuit has only m2 inductance values. For the most common machine representation with 2 field structure windings, Eq. (9-3a) requires 6 values compared to 4 values for the star circuit. This means that the star circuit de not as general as Eq. (9-3a), but this is often a blessing in disguise because available test or design data is usually not sufficient anyhow to determine all self and mtual inductances (see requirement of obtaining an extra inductance value X, in Section 8.2). As already discussed for the synchronous machine in Section 8.2, the resistances and self and mitual inductances (or the star branch inductances here) are usually not available from calculations or measurements. If the universal machine 1s used to model a synchronous machine, then the data conversion discussed in Section 8.2 can be used (input identical to synchronous machine model in version M32 and later). For three-phase induction machines, the data may be given in phase quantities. If so, Eq. (8-11) must be used to convert them to 4,q,0-quantities, Bat hg 7 be Me Loh, + mM, with L, = self inductance of one armature winding, M, = mitual inductance between two armature windings (BPA Rule Book uses opposite sign for 4,)- 1, in Bq. (8-11) is zero for an induction machine, where the saliency term defined in Eq. (8.5) does not exist. The same conversion is used if the rotor windings are three-phase. The mitual inductance between stator and rotor follows from Eq. (8.10). a Myo 7M api" y MacDi (same for q-axis), with M,_p,cosp being the mitual inductance between armature winding 1 and rotor winding Di (i=1, ... m), as defined in Eq. 9-8 (8.5). Note that the factor /3//Z changes the turns ratio; if the turns ratio between phase 1 and the rotor winding is 1:1, it changes to /3:/Z in 4,q,0-quantities (see also Section 8.2). This extra factor mst be taken into account when rotor quantities are referred to the stator side. For modelling three-phase induction machines, a modified universal machine with its own data conversion routine has recently been developed by Ontario Hydro [138]. It uses the standard NEMA specification data to find the resistances and self and mutual inductances of the equivalent circuit. It 4s beyond the scope of this treatise to describe the conversion routine in detail. Essentially, the field structure (which is the rotor in the induc- tion machine) has two windings to represent the rotor bars as well as the eddy currents in the deep rotor bars of large machines, or the double squirrel cage rotor in snaller machines. Since there is no iency, d- and qraxis paraneters are identical. The assumption of equal mitual inductances (or the star circuit) is dropped, and the equivalent circuit of Fig. 9.2 is used instead. Not surprisingly, this equivalent circuit 1s identical with Fig. 9.2 - Equivalent circuit of induction machine with deep rotor bars that of the synchronous machine in Fig. 8.23, because a synchronous machine becomes an induction machine if the field winding 1s shorted. In contrast to the standard universal machine, saturation is included in the leakage 9-9 INPUT POWER yor simution (a) active power input 7 asungnc, PHASE CURRENT (2 or summation nuns 1.2.0 (b) phase current (RMS values) “Seon! q Bod 52 a Es 8, fie) elle) (c) terminal voltage ? (RMS values) Fig. 9.3 - Comparison between field test and simulation results for starting up an induction motor with a heat transfer pump (136, 139]. Reprinted by permission of G.J. Rogers and D. Shirmohammadi 9-10 inductance branch of the armature as well, and another nonlinear inductance is added between the star point and the star branches of the field structure windings. Fig. 9.3 shows comparisons between measurements and simulation results with this modified universal machine model [138,139], for a case of a cold start-up of an induction-motor-driven heat transfer pump (1100 hp, 6600 V). Excellent agreement with the field test results is evident for the whole start-up period, which proves the validity of the modified universal machine model over the whole range of operation. 9.3 Trai ‘ormation to Phase Quantities Eq- (9-1) to (9.3) completely describe the universal machine in 4,q,0-quantities, irrespective of which type of machine it is. To solve these machine equations together with the rest of the network, they aust be transformed to phase quantities. It is in this transformation where the various types of machines must be treated differently. Fortunately it is possible to work with the same transformation matrix for all types, by simply using different matrix coefficients. For the case of a three-phase synchronous machine, the transformation has already been shown in Eq. (8.7). If this equation is rewritten for the armature quantities only, then” a 7] wf try? Jag}, tdenticat for (vl, [1] (9.7a) Xoa| with *)tn [137] and [139], (T)~! is called [T]; similarly, [P) and [S] are called (P]"! and [3]! there. oun %osp cos(B-120°) cos(p+120° A lyvz < tt) = 2 |sing sin(p-120°)sin(p+120°) (9.70) vad 1 1 1 vz vz vz being an orthogonal matrix, which means that a {T] = (7) cransposed * ene The rotor position 8 is related to the angular speed w of the rotor by w = dp/at. 9.74) The transformation matrix [T]~! can be rewritten as a product of two matrices 0137], 1 1,6)-1 m7 = etsy, (9.8a) with Tose sing tee)? = [sing —cosp ° (9.8) 0 oO L ol. lak vw eo ts)? =| 0 ee z (9.8c) v= Y= 1 1 1 ws res v3 each being orthogonal again, 1 1 ls transposed” {s] = [8] transposed G23) ‘The first transformation with the [S]~!-matrix replaces the three-phase coils (displaced by 120° in space) by the three equivalent coils d and q (perpendicular to each other) and 0 (independent by itself). This is the same transformation matrix used for a,B,0-components in Eq. (4.48), except for a sign reversal of the f-quantities. The second transformation with (P]7} makes the d,q-axes rotate with the same speed as the field poles, so 9-12 that they become stationary when seen from the field structure. ‘The field structure quantities are not transformed at all. This approach with two transformations can be applied to any type of machine. For a three-phase induction machine with a three-phase wound rotor, both the armature and field structure quantities are transformed with [S]~} to get equivalent windings on the d- and q-axes, while the transformation with [P]~} 4s only applied to the armature side. For direct current machines, there is no transformation at all for both the armature and field structure side. For two-phase armature windings displaced by 90° in space, the windings are already on the d,q-axes. Therefore 0 | 3.98) “ (= ~sing - » (9-90) sing cosp and with the zero sequence winding missing. For single-phase armature windings, there 1s only flux along one axis, or =1 (9-10a) and = cosp (9-10b) with both the quadrature axis and zero sequence winding missing. ‘The EMTP uses only one transformation matrix (S]~) and [P]~} for all cases, and makes the distinction by resetting the coefficients c,, c,, ¢, in these matrices, eytegte /Z7F —eg/o ts} = ° eglE (9-11) eglF eho 9-13 ¢3 = 1 for three-phase ac windings, and c, = 0 otherwise, 1 for two-phase ac windings, and c, = 0 otherwise, 2 1 7 1 for single-phase ac windings and de machines, and c, = 0 ° ' otherwise. Since [S]~} in Eq. (9.11) degenerates into 2 x 2 and 1x 1 matrices for two-phase and single-phase windings, its inverse cannot be found by inversion. Using Eq. (9.84) instead of inversion works in all cases, however. The matrix in Eq. (9-11) is slightly different from that in [137], because it is assumed here that only phases 1, 2 exist for two-phase machines, and only phase 1 exists for single-phase machines. In [137], phase 1 48 dropped for two-phase machines, and phases 1 and 2 are dropped for single-phase machines. For ac armature windings, [P]~! of Eq. (9.8b) is used, realizing that the zero sequence does not exist in the two-phase case, and that the zero sequence as well as the q-winding do not exist in the single-phase case. For de armature windings, there is no second transformation with (P]~!. 9-4 Mechanical Part In contrast to the synchronous machine model, the universal machine does not have a built-in model for the mechanical part. Instead, the user must convert the mechanical part into an equivalent electric network with lumped R,L,C, which is then treated by the EMTP as 1f it were part of the overall electric network. ‘The electromagnetic torque of the universal machine appears as a current source injection into the equivalent electric network. Table 9.1 describes the equivalence between mechanical and electrical quantities. For each mass on the shaft system, a node is created in the 9-14 Table 9-1 Equivalence between mechanical and electrical quantities Mechanical Electrical T (torque acting on mass) (mj| 4 (current into node) [a] w (angular speed) [rad/s]] v (node voltage) Vv] © (angular position of mass) —[rad]| vat [vs] J (moment of inertia) [kgm?]] © (capacitance to ground) — [F] K (stiffness coefficient or A/L (reciprocal of inductance) [1/H] spring constant) [Nn/rad] D (damping coefficient) [Nas/rad]|1/R (conductance) {s} (1 Nm = 0.73756 Lb-ft; Lkgm? = 23.73 1b-£t2) equivalent electric network, with a capacitor to ground with value J for the moment of inertia. If there is damping proportional to speed on this mass, a resistor with conductance D is put in parallel with the capacitor (D, in Eq. (8-31))- If there is a mechanical torque acting on that mass (turbine torque on generators, mechanical load on motors), a current source is connected to that node (positive for turbine torque, negative for load torque). If there are two or more masses, inductors are used to connect adjacent shunt capacitors, with their inductance values being equal to 1/K (reciprocal of stiffness coefficient of the shaft coupling between two masses). If there is damping ascociated with this shaft coupling, then the inductor is paralleled with a resistor whose conductance value {s D (Dj , in Eq. (8-31)). The electromagnetic torque is automatically added to the proper node as a current source by the ENTP. Fig. 9.4 summarizes the equivalence between the mechanical and electric components. Representing the mechanical system by an equivalent electric network can provide more flexibility than the built-in model of the synchronous machine of Section 8. With this approach it should be easy to Incorporate gear boxes, distributed-parameter modelling of rotors, etc- ‘The EMTP further provides for up to three universal machines sharing the same mechanical system. 9-15 MECHANICAL ELECTRICAL Le, | i, | i; co |v 2, fe, dw ew re 5 28-53% seo ar) ae cr — 1 + “TIN—_~ 2 aie ¢ Fig. 9.4 - Equivalence between mechanical and electric components Steady-State Representation and Initial Conditions 9.5 The steady-state representation of the ac-type universal machine is based on the assumption that the network to which it is connected is balanced 9-16 and linear. Only positive sequence quantities are used in the initialization, and negative and zero sequence quantities are ignored if there are imbalances. The initialization procedure could obviously be extended to handle unbalanced conditions as well, along the lines discussed in Section 8, but this extension has been given low priority so far. For three-phase synchronous machine representations, any positive sequence voltage source behind any positive sequence impedance can be used, as long as it produces the desired terminal voltages and currents when solved with the rest of the network. For simplicity, a three-phase symmetrical voltage source directly at the terminals is used for the steady-state solution. If the current (or active and reactive power output) from that solution is not what the user wants, then the power flow iteration option of the EMTP can be used, which will iteratively adjust the magnitude and angle of the three-phase voltage source until the desired active and reactive power output (or some other prescribed criteria) have been achieved. Once the terminal voltages and currents are known, the rest of the electrical machine variables are initialized in the same way as described in Section 8.4.1. If the excitation system is represented by an electric network (rather than constant ve), then the EMTP performs a second ac steady-state solution for the excitation systems of all universal machines, with the field currents i, being treated as current sources Tpcos(u;t), with w, being an angular frequency which is so low that 1, 1s de for practical purposes. This trick is used because the EMTP cannot find an exact de steady-state solution at this time (the network topology for de solutions is different from that of ac steady-state solutions; inductances would have to be treated as closed switches, capacitances as open switches, etc-)- From the initialization of the electrical variables, the electromagnetic torque 7, on the mechanical side is known from Eq. (8-32b) as well. mech-gen' These torques are used as current sources 1(t) = Taach-gen®(Ygecht) in the equivalent networks which represent the mechanical systems of all universal machines, with w...4, again being an angular frequency so low that i(t) is practically de. The EMTP then performs a third ac steady-state solution for 9-17 the initialization of the mechanical system quantities. Note that this three-step initialization procedure is direct, and does not require either predictions or iterations. 9.5.2 Two-Phase Synchronous Machine Armature currents in two-phase machines with equal amplitudes and displacements of 90° produce a rotating magnetic field in the same way as symmetrical three-phase armature currents displaced by 120°. As long as this condition is met (which is the balanced or positive sequence condition for two-phase machines), the initialization is identical with the three-phase case after proper conversion to d,q,0-quantities. If the phase quantities are 4,(t) = [I] cos(u,t + a), (9.12) p(t) I] cos (ugt + @ -90°), with w, being the (synchronous) frequency of the supply network, then the 4,q,0-quantities are obtained with [S)]~! and [P)]"! from Eq. (9.9) with ww, 4g 7 [tH sin (a- 8) (9.13) 4g [1] cos (a - 6) where 6 is the angle between the position of the quadrature axis and the real axis of the ac phasor representation. Eq. (9-13) is indeed identical with Eq. (8.41) for the balanced three-phase machine, except for a factor of V3//Z there. 9.5.3 Single-Phase Synchronous Machine Converting a single-phase armature current 44(t) = [I] cos (wt + a) (9.14) into d,q,0-quantities results in 140) = FIT] sina - 6) - F IT] since + a + 5) (9-15) i= 0 9-18, with the first term being the de quantity analogous to the positive sequence effect in three-phase machines, and the second double-frequency term analogous to the negative sequence effect in Eq. (8.53) in three-phase machines. This is a mathematical expression of the fact that an oscillating magnetic field in a single-phase armature winding can be represented as the sum of a constant magnetic field rotating forward at synchronous speed (angular speed = 0 relative to field winding) and a constant magnetic field rotating backwards at synchronous speed (angular speed = 2u relative to field winding). Since only the first term in Bq. (9.15) 1s used in the initialization now, the initial conditions are not totally correct, and it may take many time steps before steady state is reached. The steady-state torque includes @ pulsating term very similar to Fig. 8.9 for the case of an unbalanced three-phase synchronous machine. As an alternative to universal machine modelling, the three-phase synchronous machine model of Section 8 could be used for single-phase machines, by keeping two armature windings open-circuited. Unfortunately, the initialization with negative sequence quantities described in Section 8.4.2 is not yet fully correct in the BPA EMTP either, as explained in the beginning of Section 8.4, though it has been implemented in an unreleased version of the UBC EMTP. 9.5.4 De nes The initialization of de machine quantities is straightforward, and follows the same procedure outlined in Section 9.5.1. In d,q,0-quantities, balanced three-phase ac quantities appear as de quantities. Therefore, there is essentially no difference between the equations of a balanced three-phase synchronous generator and a de machine. 9.5.5 Three-Phage_ Induction Machine In balanced steady-state operation, the angular speed w of the rotor (referred to the electrical side with Eq. (8.25)) differs from the angular frequency w, of the supply network by the p.u. slip s, 9-19 : (9-16) lee Mir Fig. 9.5 - Conventional equivalent circuit for steady-state behaviour of induction machines (subscript a for armature side, subscript ¥ for rotor side) The network sees the induction machine as a positive sequence impedance whose value depends on this slip s- ‘The negative and zero sequence impedances are of no interest if the initialization is limited to balanced cases. Fig. 9.5 shows the well-known equivalent circuit for the balanced steady-state behaviour of a three-phase induction machine, which can be found in many textbooks. Its impedance can easily be calculated, and with the relationship between leakage, self and mtual inductances Lg 7 hy tM aa @.17a) Met Met becomes (ago? pos “Ba + Shag = (9.176) r . wt hie This single-phase impedance is used in phases 1,2,3 for the steady-state solution, provided there is only one winding on the field structure (rotor). For the general case of m windings on the field structure, the calculation is slightly more complicated. First, let us assume that the armature currents are 9-20 i,t) = [1] cos(u,t + a), 4,(t) = [1] cos(uzt + a - 120°), 1,(e) = |1| cos (ot +a — 240°), in balanced operation. ‘Transformed to d,q,0-quantities, the currents become a 1400) + EL tnteuge + « = 6), ae) = 8 IT] cos(sut +a - 6), (9.184) 4,() = 9, which can be represented as a phasor of slip frequency sw,, projected onto the q,d-axes, 2 36 ie = Teka (9.18b) with I = |tleJ* being the (peak) phasor current in the ac network phase solution reference frame, and with the understanding that Jou,e eu (9.180) Jou,t 10 te (ye be ALL dyq-quantities vary with the slip frequency au,, and can therefore be represented as phasors in the same way as the armature currents. To obtain the impedance, the rotor currents mist first be expressed as 4 function of armature currents. Since all rotor voltages are zero, Eq- (9-2a) can be rewritten as a O=- IR) 1A) - $e (9.19) with Dy = (yg) ty + pI] (9.20) from Eq. (9.3a) (subscript for rotor or field structure quantities, and for armature quantities). Since there is no saliency in three-phase induction machines, Eq. (9-19) and (9.20) are identical for the d~ and 9-21 qraxes, except that 1, is 1, in one case, and 1, in the other case. The submatrices [L,,] and [L,,] are obtained from the matrix of Eq. (9-3) by deleting the first row; [L,,] is the first column and (L,,] the m x mmatrix of what is left. If the rotor windings are not shorted, but connected to an R-L network, then [R,] and [L,,] mst be modified to include the resistances and inductances of this connected network (for connected networks with voltage or current sources see Section 9.5.8). Since [1,] and 1, can both be represented as phasors with Eq. (9.18), the flux in Eq. (9.20) is also a phasor which, after differentiation, becomes © [a1 = ssoyth, 5 + Jou, (L,]{T,]- (9.21) ralTga Inserting this into Eq. (9-19) produces the equation which expresses the rotor currents as a function of the armature current phasor, Ty) =~ (IR) + Seog 1) seugtt, 144 (9.22) Jo obtain the direct axis rotor currents as complex phasor quantities, use In{Ig_q} on the right-hand side of Bq. (9.22), while the use of Re{Ijg} will produce the quadrature axis rotor currents. The next step in the derivation of the iapedance is the rewriting of the armature equations (9-1a) in terms of phasor quantities. Since 4 ac haa” 8 54ga? Eq- (9-la) becomes Vad ~~ Balga 7 J8¥ghgq ~ Jtthyg or with su, = w.-w from Eq. (9.16), Yaa Balga ~ Mga (9.23) With the flux from the first row of Eq. (9.3a) aa * Maatga + Marl (t,) (9.24) where [L,,] rc) 7 [lyg]", and with Eq. (9.22), Bq. (9.23) finally becomes Vga 7 (Ry + $0, ,,) - Jo, {IRI + $80, fLyp]} "30, (L, I }Egg- Therefore, the positive sequence impedance is 9-22 2 ‘pos IE there 1s only one winding on the rotor, then it can easily be shown that the impedance of Eq. (9-17b) is identical with that of Eq. (9.25), by using the definitions of Eq. (9-17a). To summarize: The three-phase induction machine is represented as R, + July, ~ Jutb, IR] + Jsu(h, I) eugtL,.}- (9.25) three single-phase impedances Z,,, from Eq. (9.25) in the three phases 1,2,3. After the ac network solution of the complete network, the armature currents are initialized with Eq. (9-18b), and the rotor currents with Eq. (9-22). The calculation with Eq. (9-22) is done twice, with the imaginary part of I, to obtain the direct axis quantities, and with the real part of T,, to obtain the quadrature axis quantities. As mentioned before, the initialization works only properly for balanced cases at this time. If initialization for unbalanced cases is to be added some day, then the procedures of Section 8.4.2 and 8.4.3 for the synchronous machines should be directly applicable, because negative and zero sequence currents see the field winding as short-circuits. Therefore, there is no difference between synchronous and induction machines in the negative and zero sequence initialization. 9.5.6 As already discussed in Section 9.5.2 for the two-phase synchronous machine, the equations for balanced operation of a two-phase machine are identical on the d,q-axes with those of the three-phase machine. The only difference is the missing factor /3//Z in the conversion from phase quantities to d,q-quantities. 9.5.7 Single-Phase Induction Machine The problem is essentially the same as discussed in Section 9.5.3 for the synchronous machine. Only positive sequence values are used now, and the second term in age) =} [x] sin(sujt + @ -6)- 2 |r] sin((u, + we tat 5) (9.26) 2 s 2 9-23 is presently ignored. 9.5.8 Doubly-Fed_ Induct If the rotor (field structure) windings are connected to an external network with ac voltage and/or current sources, then the EMTP will automatically assume that their frequency is equal to the specified slip frequency sw, and ignore the frequency values given for these sources. Feeding the rotor windings from sources requires two modifications to the procedure of Section 9.5.5. In these modifications, it is assumed that the external network is represented by a Thevenin equivalent circuit, with voltage sources [Vp] behind an impedance matrix [Z,,,] defined at slip frequency. First, the rotor impedance matrix [R,]+Jsu,{L,,] mist be modified to Anclude the external impedances, mod (9-27) (R,] + jsw, (L. 3 ee) + rey! This modification must be done twice, for the direct axis quantities and for the quadrature =i quantities. Since [Zpey the two axes, [24] is no longer the same on both axes. Secondly, the left-hand aide of Eq. (9-19) 1s no longer zero, but J 4s in general different for Wayayls This will change Bq. (9.22) into (2,1 =~ 22287} (I gpey] + S80, lbygl Tyg) + (9.28) Again, this calculation mist be done twice. For the direct exts, use Ia{T.q) and the direct axis values znee J and [Va,y1> for the quadrature axis Re{I,g} and quadrature axis values (27 aoa ‘and vere With these ewo modifications, the steady-state model of the induction machine 1s no longer @ passive impedance Z,,,, but becomes a three-phase voltage source [E, ] behind three single-phase impedance branches source mod 2008 - Ry a2 Weta Je, ty IE mod )-1 ee $80, [hg ]}+ (9-29) The voltage source is found by calculating the direct axis contribution, 9-24 ; a4 Eg 7 Ju,(L,J(2reo4) apey-al (9.304) and the quadrature axis contribution, E. 9. 30b. 4 (9.306) and then transforming to phase quantities, 12 58 * E. ioe (Be tae) (9.30c) ‘source-1 ~ 5 ‘a ‘a H - © e75120° with Ee ource-2 “ Fsource-1" © ot . +5120" Esource-3 ~ Fsource-1 Once the ac steady-state solution of the complete network has been obtained, the d,q,0-armature currents are initialized with Eq. (9.18b), while the rotor currents are initialized with Eq. (9.28). 9.6 Transient Solution with Compensation Method For the transient solution with the compensation method, the machine differential equations (9.1) and (9.2) are first converted to difference equations with the trapezoidal rule of integration, Then Eq. (9.1) becomes vt) R, 0 OT figte gle) ] Fwceragcey] fhisey git) == [0 Ry OY fagcer | ~ Fe fagcer [+]rwcerngcey|+|nise, %oat) OO Ral Hig (€) Aga(t)]} | 0 isto, (9.312) with the history terms known from the preceding time step, 9-25 1tq Fi (eae) 8 OFLA Ce=at) g(t~At) nist, | = - |vgce-aey | - Jo 0 |fegcenaey | +2 Irgcenaey | + ae | 4 histo, Voact-at)} fo 0g [I49, ce-ae Noa (t~At) Put t=Ae) a (emt Prace-aeyag(e-ae) (9.310) 0 The field structure equations (9.2) on the direct axis become Yor (t 1 pict OT] Pet : -- te : f-2yo: fete (9.32a) ¥ pg Ct Ral }tpaCt?} OF Papg6e| [hts pa with the known history terms st yy Wp, (tae) DL pp (trae (erat fi =-| a a5 8 oe (9.320) . 7 at hist] Vp (tA) Boal Ap, (tat) Apo tA) On the quadrature axis, they are identical in form to Eq. (9.32), except that subscripts Dl, ... Dm mst be replaced by Ql, --. Qn. Finally for Eq. (9-20), 2 Vog’t? * ~ Roptogt) ~ <= Rog + M8t ops (9.334) with . - 2 - hat gg 7 Vag (EMBED ~ Roe gg(E-AE) + Rog lEBE) + (9.336) As explained in Section 12.1.2, the network connected to the armature side of the machine can be represented by the instantaneous Thevenin equivalent circuit equation 9-26 V(t 6 yee ¥26t)| = [va0} * |Requiy] ft2(#)] > (9.34) vg(t) V3.0) 15¢e) with a sign reversal for the current compared to Section 12.1.2, to change from a load to source convention. Similarly, if external networks are connected to the field structure windings, they will also be represented by Thevenin equivalent circuits with equations of the form ¥py ¢t 1-0] pit) el “Ti 1 * PRoequiv]| ! (9-35a) ©] |vy-o0 Sy C) Maa ft Yqi-o| ‘ut : =] | + [Regul] | (9.35b) Yon’ — ¥qa-o tag tt and Yoe’t) * Yor-o + Rot-equivt ot (9.356) The external network connected to the first three field structure windings is represented by a three-phase Thevenin equivalent circuit (Section 12.1.2.3), whereas the external networks connected to the rest of the field structure windings are represented by single-phase Thevenin equivalent circuits (Section 12.1.2.1). This limitation results from the fact that the BPA EMTP could handle M-phase Thevenin equivalent circuits only for M< 3 at the time the Universal Machine was first implemented. In practice, this limitation should not cause any problems because the field structure windings are usually connected to separate external networks. An exception is the three-phase wound rotor of induction machines, which is the reason why @ three-phase equivalent circuit was chosen for the first three rotor windings. The solution of the machine equations is then roughly as follows: (1) Solve the complete network without the universal machines. Extract from

You might also like