Transmission Line Grounding Volume 1 EPRI-EL—2699-Vol.1 EL-2699, Volume 1 Research Project 1494-1 DE83 900569 Final Report, October 1982 Prepared by SAFE ENGINEERING SERVICES LTD. 12201 Letellier Montreal, Quebec, Canada H3M 2Z9 Principal Author F. Dawalibi Prepared for Electric Power Research Institute 3412 Hillview Avenue Palo Alto, California 94304 EPRI Project Manager J. Dunlap Overhead Transmission Lines Program Electrical Systems Division MASTER anroimo# n> this mcukht » miMtrts DISCLAIMER This report was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof. D IS C L A IM E R Portions of this document may be illegible in electronic image products. Images are produced from the best available original document. ORDERING INFORMATION Requests for copies of this report should be directed to Research Reports Center (RRC), Box 50490, Palo Alto, CA 94303, (415) 965-4081. There is no charge for reports requested by EPRI member utilities and affiliates, contributing nonmembers, U.S. utility associations, U.S. government agencies (federal, state, and local), media, and foreign organizations with which EPRI has an information exchange agreement. On request, RRC will send a catalog of EPRI reports. Copyright © 1982 Electric Power Research Institute, Inc. All rights reserved. NOTICE This report was prepared by the organization(s) named below as an account of work sponsored by the Electric Power Research Institute, Inc. (EPRI). Neither EPRI, members of EPRI, the organization(s) named below, nor any person acting on behalf of any of them: (a) makes any warranty, express or implied, with respect to the use of any information, apparatus, method, or process disclosed in this report or that such use may not infringe private­ ly owned rights; or (b) assumes any liabilities with respect to the use of, or for damages resulting from the use of, any information, apparatus, method, or process disclosed in this report. Services Ltd. , Canada ISfC Tf’*' .|fe ABSTRACT A generalized approach to transmission line grounding has been developed and is presented in this report. Dedicated interactive computer programs based on advanced analytical models presented in the report have been developed and verified. Theoretical predictions from the computer programs are in good agreement with several measurements including data from transmission line staged fault tests conducted within the scope of this project. Typical problems are used to describe in detail the steps required to design transmission line grounds using the computer programs and design charts developed for this project. Measurement equipment and techniques are described in depth. Recommendations are provided to enhance the practicality, accuracy and usefulness of the transmission line grounding measurements. EPRI PERSPECTIVE PROJECT DESCRIPTION While considerable information is available on the design and analysis of substation grounding, relatively little is known on an equally important design consideration "Transmission Line Structure Grounding". Recent systems disturbances that may have been related to structure grounding problems emphasize a need for better design methods. As fault currents continue to grow, engineers need more accurate design methods that will assure continued safe conditions near the structures in the case of a phase-to-ground fault. This project (RP 1494-1) was initiated in order to formulate a state-of-the-art transmission structure grounding design manual that would satisfy these needs. The final report is in two volumes. Volume 1 describes analytic methods, measurement techniques, and design methodology. Volume 2 contains design charts for typical structure grounds. PROJECT OBJECTIVES The four objectives of this project were: • To assemble comprehensive background information and to develop a design methodology that would result in a complete understanding of structure grounding problems and solutions. • To provide user-oriented graphic design techniques for typically occuring situations. « To provide versatile computer programs capable of handling complex conditions. 0 To verify the accuracy of these design methods by comparing them with field test data. v PROJECT RESULTS The above objectives were fully achieved, resulting in design information on transmission line structure grounding that is superior to any previously available. This final report was specially designed for both (1) the user requiring an in-depth analysis of complex or unusual grounding problems and (2) the user wanting a quick, easy solution. The report provides sufficient information to aid the reader in handling the most complex situations. Yet, for typical designs, the charts and interactive computer program can be used by persons having little previous experience in this technology. John Dunlap, Project Manager Electrical Systems Division EPRI ACKWOWLEDeEIVIEWTS The following members of the EPRI Advisory Task Force are gratefully acknowledged for their invaluable assistance and guidance throughout the project: J. Dunlap, Project Manager, Electric Power Research Institute T. E. Bethke, Potomac Electric Power A. C. Pfitzer, Tennessee Valley Authority C. J. Blattner, Niagara Mohawk G. B. Niles, Baltimore Gas and Electric R. S. Baishiki, Pacific Gas and Electric Mr. G. B. Niles is particularly thanked for his assistance in the testing and evaluation of the software package GATE. Especially acknowledged is the excellent cooperation provided by the Tennessee Valley Authority (TVA) and Rochester Gas and Electric Company (RGE) who participated in the planning and execution of the field test programs. Special thanks are extended to Mr. J. W. Chadwick from TVA and Mr. J. Windsor from RGE. Appreciation is extended to Electricite de France, Ontario-Hydro, American Electric Power Company and Idaho Power Company for providing much data valuable for this project. The following individuals are particularly thanked for their willingness to share information and discuss subjects related to this project: P. Kouteynikoff, Electricite de France; E. A. Cherney, Ontario Hydro; N. Kolcio, American Electric Power; W. G. Eisinger, Idaho Power. Deserving of special acknowledgment for services and assistance provided during the project is Mr. D. Bensted from David Bensted and Associates Ltd.. Mr. Bensted is the principal author of Chapter 9 and was responsible for the coordination and execution of the field tests. All of the utilities that responded to the transmission line grounding survey questionnaire are gratefully acknowledged for their invaluable contributions. CONTENTS Page Chapter 1 2 INTRODUCTION 1.1 GENERAL 1-1 1.2 HISTORIC NOTES 1-1 1.3 PROJECT OBJECTIVES 1-2 1.4 USE OF THIS BOOK 1-3 1.5 CONCLUDING REMARKS 1-5 FUNDAMENTAL CONSIDERATIONS 2.1 GENERAL 2-1 2.2 TRANSMISSION LINE DISTURBANCES 2-1 2.3 EARTH 2.3.1 Simplified Methods 2.3.2 Computer Program RESIST 2-2 2-3 2-3 2.4 POWER SYSTEM 2.4.1 Conductor 2.4.2 Computer 2.4.3 Computer 2-3 2-5 2-7 2-7 2.5 GROUNDING PERFORMANCE OF TRANSMISSION LINE STRUCTURES 2.5.1 Simplified Formulas 2.5.2 Computer Program GTOWER 2-8 2-8 2-8 2.6 TYPE OF DISTURBANCE 2.6.1 Power Frequency Ground Faults 2.6.2 Lightning Strokes 2-9 2-9 2-9 NETWORK Impedances Program LIN PA Program PATHS REFERENCES 2-10 ix CONTENTS (continued) Page Chapter 3 4 TYPICAL DESIGN METHODOLOGY 3.1 GENERAL 3-1 3.2 DESCRIPTION OF THE PROBLEM 3.2.1 Tower Footing 3.2.2 Power System Network 3.2.3 Transmission Line Conductor Data 3.2.4 Apparent Soil Resistivity 3-1 3-2 3-2 3-4 3-5 3.3 DESIGN METHODOLOGY 3.3.1 Two Layer Earth Model 3.3.2 Transmission Line paramaters 3.3.3 Tower Grounding Performance 3.3.4 Fault Current Distribution 3.3.5 Safety Evaluation 3-5 3-6 3-6 3-9 3-13 3-14 3.4 POTENTIAL CONTROL CONDUCTORS 3.4.1 Inner Potential Control Loop; Step 1 3.4.2 Leg Control Loop; Step 2 3-15 3-15 3-17 3.5 OTHER CASES 3-18 EARTH RESISTIVITY 4.1 GENERAL 4-1 4.2 ELECTRICAL PROPERTIES OF EARTH 4.2.1 Conduction by Electronic Process 4.2.2 Electrolytic Conduction 4.2.3 Electrical Conduction in Earth 4.2.4 Electrical Model of Earth 4-2 4-3 4-3 4-5 4-8 4.3 MEASUREMENT METHODS AND TECHNIQUES 4.3.1 Electrical Well Logging 4.3.2 Galvanic Resistivity Methods 4.3.3 The Wenner Method 4-8 4-9 4-11 4-12 4.4 DERIVATION OF THE EQUATIONS 4.4.1 Point Source Electrode in Uniform Soil 4.4.2 Multiple Source Electrodes 4.4.3 Nonuniform Soils 4.4.4 Point Source Electrode in a Two-Layer Earth 4.4.5 Point Source in Earth with Inclined Layers 4.4.6 Point Source Embedded in a Localized Discontinuity 4.4.7 Application of the Potential Functions to the Wenner Array 4-14 4-14 4-17 4-18 4-19 4-22 4-23 4-24 x CONTENTS (continued) Chapter 4.5 Page INTERPRETATION OF THE MEASUREMENTS 4.5.1 Basic Considerations 4.5.2 Empirical Methods 4.5.3 Analytical Methods 4.5.4 Methodology 4.5.5 Logarithmic Curve Matching 4.5.6 Partial Curve Matching 4.5.7 Other Methods 4.5.8 The Steepest Descent Method REFERENCES 5 4-27 4-27 4-28 4-30 4-32 4-33 4-35 4-38 4-39 4-42 POWER FREQUENCY PERFORMANCE OF TRANSMISSION LINE STRUCTURE GROUNDS 5.1 GENERAL 5-1 5.2 TRANSMISSION LINE GROUNDING 5.2.1 Transmission Line Structure Grounding System 5.2.2 Equivalent Cylindrical Conductor 5-3 5-3 5-5 5.3 LOW FREQUENCY RESPONSE OF STRUCTURE GROUNDING SYSTEMS 5.3.1 Grounding Performance of a Line Source 5.3.2 Analysis of a Structure Grounding System 5-7 5-7 5-12 5.4 LOCAL HETEROGENEITIES IN THE SOIL 5.4.1 Hemispherical Electrode Embedded in a Hemispherical Shell 5.4.2 Vertical Rod Embedded in a Cylindrical Shell 5.4.3 Vertical Rod Encased in an Elliptic Shell 5.4.4 Comparison Between the Equipotential Surface Methods 5.4.5 Extensions & Limitations of the Equipotential Surface Methods 5.4.6 Concrete Encased Grounding Elements 5-18 5-18 5-21 5-23 5-27 5-29 5-32 5.5 CONTINUOUS COUNTERPOISES 5.5.1 Performance of an Isolated Continuous Counterpoise 5.5.2 Counterpoise of Transmission Lines 5.5.3 Accurate Analysis of Counterpoises 5-35 5-36 5-41 5-47 5.6 SOIL HEATING EFFECTS 5.6.1 Heating of the Soil Around a Grounding System 5.6.2 Steady State Performance 5.6.3 Transient Heating 5-50 5-50 5-51 5-52 REFERENCES 5-54 xi CONTENTS (continued) Chapter 6 Page POWER FREQUENCY FAULT CURRENT DISTRIBUTION 6.1 GENERAL 6-1 6.2 IMPORTANCE OF FAULT CURRENTDISTRIBUTION 6-2 6.3 CONSTANT LINE PARAMETERS 6.3.1 Long Transmission Lines 6.3.2 Long and Short Transmission Lines 6-3 6-3 6-6 6.4 VARYING LINE PARAMETERS 6.4.1 Sebo's Method 6.4.2 The Single and Double Sided Elimination Methods 6.4.3 The Generalized Double Sided EliminationMethod 6.4.4 Practical Considerations 6-12 6-13 6-16 6-23 6-30 REFERENCES 7 6-33 LIGHTNING PERFORMANCE OF TRANSMISSION LINE STUCTURES 7.1 GENERAL 7-1 7.2 EFFECTS OF LIGHTNING ON POWER LINES 7.2.1 The Lightning Flash 7.2.2 The Back Flashover Mechanism 7-1 7-1 7-2 7.3 ANALYTICAL MODELS FOR BACK FLASHOVERCOMPUTATIONS 7.3.1 Simplified Models 7.3.2 Travelling Wave Models 7.3.3 Electromagnetic Field Models 7-5 7-5 7-8 7-13 7.4 UNCERTAINTIES IN THE ANALYTICAL MODEL OF A STRUCTURE 7.4.1 Lossless Conducting Elements 7.4.2 Earth Resistivity 7.4.3 Analytical Models of Extended Grounding Systems 7.4.4 Soil Ionization Mechanism 7-19 7-19 7-20 7-21 7-23 7.5 FREQUENCY DOMAIN APPROACH 7.5.1 Frequency Effects on Electrical Constants of Materials 7.5.2 Clarity and Precision 7.5.3 Hybrid Models 7.5.4 Conclusion 7-25 7-25 7-26 7-27 7-27 REFERENCES 7-27 xii CONTENTS (continued) Chapter 8 Page SAFETY 8.1 GENERAL 8-1 8.2 THE ELECTROCUTION EQUATION 8.2.1 The Electrocution Mechanism 8.2.2 Thresholds of Current 8.2.3 Parameters Influencing Heart Fibrillation 8.2.4 The Electrocution Theories 8.2.5 The Electric Shock 8.2.6 Safe Impulse or Surge Current 8-2 8-3 8-3 8-3 8-7 8-8 8-14 8.3 PROBABILISTIC CONSIDERATIONS 8.3.1 The Probability of an ElectrocutionIncident 8.3.2 The Statistical Probabilistic Approach 8.3.3 The Selective Deterministic Approach 8-16 8-16 8-16 8-19 8.4 SAFE DESIGNS AND MITIGATION TECHNIQUES 8.4.1 Evaluation of the Safety Level 8.4.2 Improving Safety Around TransmissionLine Structures 8.4.3 Ground Potential Mitigating Conductors 8.4.4 Other Safety Measures REFERENCES 9 8-20 8-21 8-22 8-25 8-27 8-27 FIELD MEASUREMENT TECHNIQUES 9.1 GROUND RESISTANCE MEASUREMENT 9-1 9.2 TEST METHOD 9.2.1 Two-Point method 9.2.2 Three-Point Method 9.2.3 Fall-of-Potential Method 9.2.4 Theory of the Fall-of-Potential Method 9.2.5 Interpretation of Fall-of-Potential Data 9.2.6 Overhead Ground Wires 9-2 9-2 9-2 9-3 9-4 9-6 9-9 9.3 GROUND RESISTANCE MEASUREMENT INSTRUMENTATION 9.3.1 Portable Ground Testers 9.3.2 Voltmeter/Ammeter Method 9.3.3 High Frequency Portable Instruments 9.3.4 Other Measurement Systems 9-10 9-10 9-11 9-12 9-12 xi i i CONTENTS (continued) Chapter Page 9.4 MEASUREMENT TECHNIQUES 9-13 9.5 EARTH RESISTIVITY TESTS 9.5.1 The Four Point Method 9.5.2 Driven Ground Rod Method 9.5.3 Resistivity MeasurementInstrumentation 9.5.4 Measurement Techniques 9-16 9-17 9-18 9-19 9-20 9.6 SAFETY PRECAUTIONS 9-21 REFERENCES 10 9-22 COMPARISON BETWEEN MEASUREMENTS AND COMPUTATIONS 10.1 GENERAL 10-1 10.2 THE RGE 115 kV TRANSMISSION LINE MEASUREMENTS 10.2.1 Test Site 10.2.2 Test Equipment 10.2.3 Earth Resistivity Tests 10.2.4 Analysis of the Resistivity Tests 10.2.5 Ground Impedance Measurements 10.2.6 Analysis of the Ground Impedance Measurements 10-2 10-2 10-5 10-6 10-7 10-11 10-13 10.3 THE TVA 500 kV TRANSMISSION LINE MEASUREMENTS 10.3.1 Test site 10.3.2 500 kV Staged Fault Test Description 10.3.3 Low Current Tests 10.3.4 SES Test Procedure 10.3.5 Earth Resistivity Tests 10.3.6 Analysis of the Resistivity Tests 10.3.7 Earth Potential and Ground Impedance Tests 10.3.8 Analysis of the High Frequency Tests 10.3.9 Analysis of the Low Frequency Tests 10.3.10 Analysis of the TVA Staged Fault Current Measurements 10-16 10-19 10-20 10-22 10-23 10-24 10-28 10-31 10-33 10-35 10-42 10.4 THE OH/AEP 500/765 kV TRANSMISSION LINE MEASUREMENTS 10.4.1 The AEP Measurements 10.4.2 The OH Measurements 10-50 10-50 10-54 10.5 THE EDF ROCKET TRIGGERED LIGHTNING MEASUREMENTS 10.5.1 Equivalent Circuit of the Tower Structure 10.5.2 Equivalent Circuit of the Grounding Electrodes 10-59 10-62 10-63 REFERENCES 10-64 xiv CONTENTS (continued) Chapter 11 Page SURVEY OF NORTH AMERICAN TRANSMISSION LINE GROUNDING PRACTICES 11.1 OBJECTIVES OF SURVEY 11-1 11.2 SURVEY DISTRIBUTION 11-1 11.3 RESPONSE TO THE SURVEY 11-2 11.4 QUESTIONNAIRE FORMAT 11-2 11.5 ANALYSIS OF THE RESPONSES 11.5.1 General 11.5.2 Measurements 11.5.3 Engineering 11.5.4 Safety Considerations 11.5.5 Construction/Measurements/Maintenance 11.5.6 Typical Design and Data 11-3 11-3 11-4 11-4 11-5 11-6 11-7 11.6 SUMMARY OF COMMENTS RECEIVED 11-11 11.7 CONCLUSION 11-14 APPENDIXES A COMPUTER PROGRAM LIN PA A-l B COMPUTER PROGRAM RESIST B-l C COMPUTER PROGRAM GTOWER C-l D COMPUTER PROGRAM PATHS D-l E LIST OF SURVEY RESPONDENTS E-l F RGE 115 KV TRANSMISSION LINE GROUNDING TEST RESULTS F-l G TVA 500 KV STAGED FAULT AND TOWER GROUNDING TEST RESULTS G-l ILLUSTR AXIOMS Page Figure 2.1 Equivalent Power System Network 2-4 3.1 345 kV Four Leg Lattice Tower 3-2 3.2 Grillage Foundations 3-3 3.3 Power System Network 3-3 3.4 Design Methodology 3-7 3.5 Logarithmic Curve Matching Technique 3-8 3.6 Worst Touch and Step Voltages 3-9 3.7 Fundamental Tower Grounding System Dimensions 3-11 3.8 Earth Surface Potential Profiles 3-12 3.9 Simplified Circuit 3-13 3.10 Ground Potential Control Loops 3-16 4.1 Soil Resistivity Versus Temperature 4-5 4.2 Soil Resistivity Versus Temperature for Typical Soil Samples 4-6 4.3 Resistivity Map of the United States 4-7 4.4 Geoelectric Model of Earth 4-8 4.5 Operating Principle 4-10 4.6 Commonly Used Arrangements 4-10 4.7 Typical Log Resistivity Curve 4-11 4.8 Galvanic Resistivity Methods 4-12 4.9 Wenner Arrangement 4-13 4.10 Point Source Electrode in Uniform Soil 4-15 4.11 Hemispherical Electrode 4-16 4.12 Schlumberger-Palmer Array 4-18 4.13 Two-Layer Earth Model 4-19 xv ii ILLUSTRATIONS (continued) Page Figure 4.14 Vertical Fault 4-22 4.15 Localized Discontinuity 4-24 4.16 Wenner Array 4-25 4.17 Vertical Fault (Top View) 4-25 4.18 Uniform Earth 4-27 4.19 Sudden Changes 4-28 4.20 Current Density in Earth 4-29 4.21 Assymptote's Rules 4-31 4.22 Logarithmic Curve Matching 4-34 4.23 Horizontal Two-Layer Earth Master Chart 4-35 4.24 Vertical Fault Master Chart 4-36 4.25 Multi-Layer Earth 4-37 4.26 Partial Curve Matching 4-37 4.27 The Method of Steepest-Descent 4-39 5.1 Equivalent Cylindrical Conductor 5-5 5.2 Equivalent Linear Current Source 5-7 5.3 Point Source Below Earth Surface 5-8 5.4 Infinite Number of Point Sources 5-9 5.5 Segmentation Process 5-10 5.6 Tower Grounding System 5-13 5.7 Average Potential and Center Methods 5-17 5.8 Local Soil Heterogeneities 5-18 5.9 Hemispherical Configurations 5-19 xv i i i ILLUSTRATIONS (continued) Page Figure 5.10 Effects of a Hemispherical Heterogeneity 5-20 5.11 Tower Foundations 5-20 5.12 Small Cylindrical Heterogeneous Material 5-21 5.13 Effects of a Cylindrical Heterogeneity 5-23 5.14 Elliptic Shell 5-24 5.15 Ellipsoid 5-24 5.16 Effects of an Elliptic Heterogeneity 5-28 5.17 Equal Volume Base 5-28 5.18 The Equipotential Metallic Sheet Principle 5-29 5.19 Illustration of a Typical Problem 5-30 5.20 Typical Concrete Foundations 5-32 5.21 Dynamic Resistance of Concrete 5-33 5.22 Arcing Process in a Concrete Distribution Pole 5-35 5.23 Isolated Continuous Counterpoise 5-36 5.24 Infinitesimal Counterpoise Element 5-36 5.25 Ground Impedance of a Counterpoise 5-41 5.26 Effects of Frequency on Counterpoise Impedance 5-42 5.27 Transmission Line Counterpoise 5-43 5.28 Equivalent Conductor Pair for a Counterpoise 5-44 5.29 First Iteration 5-45 5.30 Last Iteration 5-46 5.31 Computer Solution of Counterpoise Ground Impedance 5-48 5.32 Current Distribution Along a Counterpoise 5-49 xix ILLUSTRATIONS (continued) Figure Page 5.33 Heating of the Soil Around a Grounding Electrode 5-51 5.34 Temperature Variation with Time 5-53 6.1 Conventional Circuit Reduction 6-2 6.2 Fault Current Distribution 6-3 6.3 Distributed Parameters Method 6-4 6.4 Lumped Parameters Method 6-7 6.5 Distributed & Lumped Method - ConstantTower Resistance 6-11 6.6 Currents in Ground Wire 6-13 6.7 Circuit of One Span of a Line 6-15 6.8 Ground Fault at Span n From the SourceTerminal 6-16 6.9 Equivalent Network - General Case 6-17 6.10 Fundamental Loops of the Network 6-18 6.11 Fundamental Loops of the Network 6-24 6.12 A Phase-to-Ground Fault on a Transmission Line 6-31 6.13 Effects of Ground Resistance Values 6-32 7.1 A Typical Transmission Line Back Flashover 7-3 7.2 Travelling Waves After a Lightning Stroke toa Structure 7-4 7.3 Simplified Models 7-6 7.4 Approximation of a Current Surge 7-7 7.5 Lightning and Travelling Waves 7-9 7.6 Illustration of a Typical Problem 7-12 7.7 Typical Structure Top Potentials 7-12 7.8 Moving Wave of Electric Charges 7-13 xx ILLUSTRATIONS (continued) Figure Page 7.9 Integration Path of Vector Potential 7-16 7.10 Voltage Across Insulator String 7-16 7.11 Wave Reflection and Refraction Technique 7-18 7.12 Earth Resistivity and Earth Potentials 7-20 7.13 Dynamic Resistivity-Impulse Current Curve 7-23 7.14 Frequency Dependence of Material ElectricalConstants 7-26 8.1 Strength-Duration Stimulus Curve 8-4 8.2 Equivalent Circuit of a Typical Cell Membrane 8-5 8.3 Threshold of Ventricular Fibrillation for a 50Kg Man 8-6 8.4 Safe Body Current Versus Time 8-8 8.5 The Electric Shock Circuit 8-9 8.6 Simplified Equivalent Electric Shock Circuit 8-11 8.7 Body Resistance Versus Applied Voltage 8-12 8.8 Ground Resistance of Feet 8-13 8.9 Nonfatal Surge Capacitor Discharge Currents 8-14 8.10 Perception Current as a Function of Frequency 8-15 8.11 Approximate Equivalent Transmission Line Circuit 8-22 8.12 Body Current Versus Resistivity 8-24 8.13 Resistivity of Gravel 8-25 8.14 Typ ical Ground Potential Mitigating Rings 8-26 9.1 Principle of Ground Resistance Measurement 9-2 9.2 Fall-of-Potential Test Schematic Diagram and Typical Test Results 9-4 ILLUSTRATIONS (continued) Page Figure 9.3 Required Potential Probe Location in Fall-of-Potential Measurements 9-6 Required Potential Probe Location in a Two-Layer Structure 9-8 9.5 F all-of-Potenti al Profile Results ofan Actual Test 9-9 9.6 Voltmeter/Ammeter Test Set-Up 9-11 9.7 Functional Schematic - High Current Frequency Selective Ground Resistance Test Set-Up 9-12 General Instrumentation Arrangement for the Measurement of Ground Impedance Using a FFTDigital Analyzer 9-13 Field Calibration Check or Ground Resistance Measuring Equipment 9-17 Four Point Method of Earth ResistivityMeasurement 9-19 9.4 9.8 9.9 9.10 10.1 Rochester Gas and Electric Test Site 10-3 10.2 115 kV Wood Structures 10-3 10.3 Curent (Cl) & 10-4 10.4 Test Equipment 10-5 10.5 Functional Schematic of the Test Set-Up 10-6 10.6 Rochester Gas and Electric Measurement Site 115 kV Transmission Line 10-7 10.7 Earth Resistivity Traverse RGE 01-1 10-8 10.8 Earth Resistivity Traverse RGE 01-2 10-9 10.9 Earth Resistivity Traverse RGE 03 10-9 Potential (PI) Test Lead Connections 10.10 Earth Resistivity Traverse RGE 04 10-10 10.11 Equivalent Circuit of Grounding Connections at Double-Pole Structures 10-13 xx ii ILLUSTRATIONS (continued) Figure Page 10.12 Frequency Dependence of Transmission Line Grounds 10-14 10.13 Measured and Computed Apparent Impedance 10-16 10.14 Test Site; Tower 130 and Fault Structure 10-17 10.15 Arc Initial at Fault Structure 10-17 10.16 Staged Fault Tests Measurements 10-19 10.17 Topography of the Test Site 10-20 10.18 Potential Transformers and Crossbow NearTower 130 10-21 10.19 Resistivity Measurements 10-23 10.20 Resistivity Traverse TVA 1-1 10-24 10.21 Resistivity Traverse TVA 2-2 10-23 10.22 Resistivity Traverses TVA 1-2 and TVA 10 10-26 10.23 Resistivity Traverses TVA 2-1 and TVA 03 10-27 10.24 Resistivity Traverse TVA 04 10-28 10.25 Potential Profile Measurements (Ground Wires Connected to Tower 130 & 10 Adjacent Towers on Each Side of Tower 130) 10-31 10.26 Potential Profile Measurements (Ground Wires Disconnected From Tower 130) 10-32 10.27 500 kV Tower Leg 10-33 10.28 Equivalent Circuit with Ground Wires Connected 10-34 10.29 Measured and ComputedGroundImpedances 10-35 10.30 A Typical TVA 500 kV Tower 10-36 10.31 500 kV Tower Leg Earth Anchor 10-37 xxiii ILLUSTRATIONS (continued) Figure Page 10.32 Equivalent Earth Anchor Grounding Model 10-37 10.33 Measured and Computed Potenti al Profiles (Ground Wires Connected to Tower) 10-38 10.34 Computed Potential Profiles (soil Models 1 to 5) (Insulated Ground Wires) 10-40 10.35 Measured and Computed Potential Profiles (Insulated Ground Wires) 10-40 10.36 Measured and Computed Potential Profiles Compatible Soil Models (Insulated Ground Wires) 10-41 10.37 Equivalent Positive and Zero-Sequence Network 10-42 10.38 Equivalent Faulted Transmission Line Circuit 10-43 10.39 Transmission Line Conductor Impedance 10-44 10.40 Current and Voltage Oscillograms Measured at Tower Site 10-47 10.41 Current and Voltage Oscillograms Measured at Cumberland Station 10-48 10.42 Current and Voltage Oscillograms Measured at Johnsonville Station 10-49 10.43 The AEP 765 kV Test Sites 10-50 10.44 Plan View of Guyed-Vee Towers 10-51 10.45 Average Apparent Soil Resistivities Around Guyed-Vee Towers 10-52 10.46 Earth Potential Profiles for Tower 13 10-53 10.47 Earth Potential Profile for Tower 36 10-53 10.48 Plan View of 765 kV Lattice Tower Footing 10-54 10.49 500 kV Tower Leg 10-54 at Kleinburg 10.50 Plan Cross Sectional View Views of the 500 kV Tower Footing XXIV 10-55 ILLUSTRATIONS (continued) Figure Page 10.51 Earth Surface Potential Measurements pads 10-55 10.52 Resistivity Measurements at Kleinburg Test Site 10-56 10.53 Earth Surface potential Profiles for Ground Rods and GPC Ring 10-57 10.54 Earth Surface Potential Profiles for Rebar Cages and GPC Ring 10-57 10.55 Earth Surface Potential Profiles for Ground Rods With Rebar Cages and GPC Ring 10-58 10.56 Tower Structure at St. Privat d'Allier 10-59 10.57 Ground Level Rocket-Launching Area 10-60 10.58 Tower Base and Insulator Supports 10-61 10.59 Electrode Configuration Subjected to Surge Tests 10-61 10.60 Tower Structure Impedance 10-62 10.61 Ground Impedance of Hemispherical, Double Loop and Horizontal Electrodes 10-63 10.62 Ground Impedance of a Ground Rod and a TwoBranch Star 10-63 11.1 Size of Utilities Which Answered the Questionnaire 11-2 11.2 Distribution of Fault Current Between Structure and Skywires 11-5 11.3 Additional Cost of Grounding 11-6 11.4 Average Useful Life of Grounding Conductors 11-7 11.5 Grounding Arrangement 11-9 11.6 Transmission Structure Resistance 11-10 11.7 Lightning Rate of Failure 11-10 F. l RGE'S 115 KV Transmission Line Test Site F-l G. l TVA'S 500 kV Johnsonville-Cumberland Transmission Line G-l TABLES Table Page 3.1 Apparent Soil Resistivity 3-3 3.2 Computed Impedance matrix 3-8 3.3 Permissible Contact Voltages 3-15 4.4 Typical Resistivity Values 4-4 6.1 Round-off Errors in Computations 6-20 6.2 Initial Values 6-22 9.1 Resistance Measurement Equipment Check List 9-14 9.2 Mutual Impedance Between Two Parallel Conductors on the Surface of the Earth 9-16 10.1 Equivalent Two-Layer Soils 10-8 10.2 Pole and Structure Ground Resistances 10-14 10.3 Frequency Response of Transmission Line Structures 10-15 10.4 Equivalent Two-Layer Soils 10-29 10.5 Origin of Soil Models Analyzed 10-30 10.6 Ground Wire Impedance 10-34 10.7 Computed Ground Resistances 10-39 10.8 Bonded Ground Wires Test Results & Computations 10-45 10.9 Insulated Ground Wires Test Results 10-46 10.10 Guyed-Vee Tower Resistances 10-52 10.11 Kleinburg Tower Ground Resistance 10-56 11.1 Maximum Tower Ground Resistances 11-8 11.2 Transmission Line Lightning Failure Rates 11-8 xx vi i TABLES (continued) Table Page F.l Resistivity Measurements - Traverse RGE 01-1 F-3 F.2 Resistivity Measurements - Traverse RGE 01-2 F-3 F.3 Resistivity Measurements - Traverse RGE 03 F-4 F.4 Resistivity Measurements - Traverse RGE 04 F-4 F.5 Potential Measurements - Profile RGE 02-1 F-5 F.6 Potential Measurements - Profile RGE 02-2 F-5 F.7 Potential Measurements - Profile RGE 02-3 F-6 F.8 Potential Measurements - Profile RGE 02-4 F-6 F.9 Potential Measurements - Profile RGE 02-4 F-7 F.10 Potential Measurements - Profile RGE 02-5 F-7 F.ll Potential Measurements - Profile RGE 02-7 F-8 F.12 Potential Measurements - Profile RGE 02-8 F-8 F.13 Potential Measurements - Profile RGE 02-9 F-9 RGE 02-10 F-9 F. 14 Potential Measurements - Profile G. l Resistivity Measurements - Traverse TV A 1-1 G-3 G.2 Resistivity Measurements - Traverse TV A 1-2 G-3 G.3 Resistivity Measurements - Traverse TV A 2-1 G-4 G.4 Resistivity Measurements - Traverse TV A 2-2 G-4 G.5 Resistivity Measurements - Traverse TV A 03 G-5 G.6 Resistivity Measurements - Traverse TV A 04 G-5 G.7 Resistivity Measurements - Traverse TV A 10 G-6 G.8 Potential Measurements - Profile TV A 01 G-7 G.9 Potential Measurements - Profile TV A 01 G-7 G.10 Potential Measurements - Profile TV A 02 G-8 xxviii TABLES (continued) Table Page G.ll Potential Measurements - Profile TV A 11 G-8 G.12 Potential Measurements - Profile TV A 05 G-9 G.13 Potential Measurements - Profile TV A 06 G-9 G.14 Potential Measurements - Profile TVA 07 G-10 G.15 Potential Measurements - Profile TVA 08 G-10 G.16 Potential Measurements - Profile TVA 09 G-ll G.17 Ground Resistance Measurements G-13 G.18 Ground Resistance Measurements G-13 G.19 Ground Resistance Measurements G-14 xxix SUMMARY OBJECTIVES The primary goal of this Electric Power Research Institute (EPRI) project has been to develop a transmission line grounding reference book which consolidates ail available information on the subject and describes a systematic and practical design methodology. In meeting this objective, two volumes have been written and a software package developed. RESULTS Volume 1 of the final report describes analytical methods, measurement techniques and design methodologies related to transmission line grounding. A summary of survey responses from North American utilities on the state-of-practices of transmission line grounding is included. Volume 1 also contains instructions for using the software package. Volume 2 contains an extensive set of design curves intended to provide an alternative design methodology to the computer analysis approach. The software package GATE (Grounding Analysis of Transmission Lines) consists of four FORTRAN programs developed to provide simple and accurate tools for the analysis and design of transmission line grounding systems and ground potential mitigation electrodes. CONTENTS A significant portion of the information contained in this report has come from published work. This information has been grouped in a manner convenient for use by transmission line engineers. New methods and techniques are also presented. Examples of original research work include earth resistivity analysis and interpretation techniques, an analysis of concrete encased electrodes, computer generated counterpoise impedance curves and transmission line fault current distribution analysis. New concepts and approaches have also been introduced in the areas of field measurement techniques and safety. S-1 VOLUME 1 The first part of the report presents practical information on transmission line grounding design techniques. A detailed description of the design data required by GATE or by the design graphs of Volume 2 is provided. Techniques for interpreting the results obtained from the programs and from the charts are described. The middle portion of the report describes advanced analytical methods relating to transmission line grounding. Recommendations for earth resistivity measurement and interpretation techniques are given. Equations that predict the power frequency performance of a transmission line grounding system are supplied. The subjects of concrete encased electrodes, counterpoises and soil heating are also discussed. Methods for calculating the current distribution between a faulted structure and the overhead ground wires are provided. The report analyzes the response of transmission line structures to lightning strokes and discusses existing analytical theories. The latter portion of the report describes modern theories on the effects of body currents and introduces the probabilistic and deterministic approaches to safety in the vicinity of transmission line structures. Measurement techniques and equipment are described. Actual field tests, including staged fault tests conducted during this EPRI project are also reported. The purpose of these tests was to provide a comparison of measured and computed results. Good agreement was obtained between field measurements and results of the calculations. VOLUME 2 Volume 2 contains design curves, derived primarily from the results of the GATE computer program. Charts permit the user to convert resistivity measurements into equivalent earth models, to design grounding systems, to determine the current distribution in a faulted line and to perform safety analysis. SOFTWARE PACKAGE GATE GATE consists of four programs: LINPA, RESIST, GTOWER and PATHS. LINPA calculates the self and mutual impedances of transmission line conductors, RESIST interprets earth resistivity measurements, GTOWER analyzes the performance of transmission line grounding systems and PATHS calculates the distribution of fault current in the transmission line. CONCLUSION The two volumes of the reference book and the GATE computer programs form an integrated package which meets the widely varying needs of researchers and designers. They also serve as a comprehensive reference document for engineers with little or no grounding experience. S-2 iWTRODUCTiOW 1.1 GENERAL This book is the culmination of several years of design and research work in the field of grounding. Much of the work was completed by Safe Engineering Services Ltd (SES) and other organizations prior to the start of the project. All available research results were combined and adapted to a transmission line design perspective in convenient reference volumes. These documents contain the information and tools necessary to conduct an engineering study that fully utilizes the capabilities of today's computer technology. 1.2 HISTORIC NOTES The traditional approach to transmission line grounding design is highly empirical. Practices vary widely from one utility to the other as evidenced by the results of a North American utility survey summarized in Chapter 11 of this report. The design of a power system reguires that both normal and abnormal conditions be considered in order to correctly determine the design reguirements and characteristics of the installed power equipment. Of the abnormal conditions which can occur on a power transmission line, the two most frequent are: • Lightning strokes • Phase to ground faults In both cases, the overhead network and the earth path, including buried metallic conductors such as counterpoises and ground electrodes, are part of the circuit in which the surge or fault current circulates. Generally, an analysis of these abnormal conditions is based on a reasonably accurate representation of the overhead circuit. The earth path, however, is usually modelled as a perfect conductor or in a very simplified form. This seldom leads to realistic results. The apparent inconsistancy of these engineering approaches can be explained by the mathematical difficulties involved in the analysis of three dimensional current flow in earth. Often, the wide variations observed in the characteristics of earth, generally described as a semi-infinite nonhomogeneous medium, are used as justification for not pursuing detailed modelling of the earth path for fault currents. Thirty years ago, the lack of suitable high-speed digital computers was a serious obstacle to accurate modelling of the earth. Now, there are no computational limitations to the development of an accurate model of the structure of the earth. Recently published analytical works on power system grounding describe accurate, computer based computational techniques for the design of grounding systems. 1-1 Large variations in earth resistivity need not be an obstacle to the development of detailed earth structure models. Relatively simple equivalent earth models can be effectively used to accurately predict transmission line grounding performance, as evidenced by the field measurements described in Chapter 10. Finally, the earth structure at any particular site can be accurately determined by a suitable selection of test methods and equipment. Tremendous increases in energy consumption coupled with environmental constraints which limit the availability of right of ways, have introduced a particular need for accurate and practical methods for the analysis and design of cost effective transmission line grounding systems and potential gradient mitigation schemes. 1.3 PROJECT OBJECTIVES The objectives of the project have been to provide to the Electric Power Industry the following items consolidated into a single, convenient packages • A reference document including a comprehensive description of advanced theories and techniques pertaining to the analysis, design and measurement of transmission line grounding systems, with a particular emphasis on safety and mitigation techniques to improve safety around exposed structures. • A field measurement report on actual transmission lines including high current staged fault tests in order to prove the effectiveness of the advanced analytical methods described in the report. • A practical and concise design manual which clearly describes the various steps of transmission line grounding system design using realistic examples. • A series of dedicated computer programs which can be used in a batch or interactive mode and require a minimum of grounding expertise on behalf of the user, while delivering a maximum of computed information based on the sophisticated theories and concepts described in the final report. • An appropriate selection of design charts for those who do not have access to a computer terminal, or wish to obtain quick estimates of transmission line grounding parameters for standard transmission line grounding systems. In achieving these challenging objectives, new methods and techniques, not previously available in the literature, have been introduced. Examples of such new developments include earth resistivity interpretation techniques, concrete encased electrode design, computer generated counterpoise impedance charts, and transmission line fault current distribution analysis by the generalized double-sided elimination method. Original contributions have also been made in areas such as safety and field measurement techniques. Much of the information contained in this report comes from published research work. This has been conveniently regrouped from a transmission line engineering viewpoint. 1-2 1.4 USE OF THIS BOOK To simplify the use of this book, it has been organized in two volumes. Volume 1 contains the bulk of the text while Volume 2 contains the design charts. Each chapter is generally self-contained and includes the necessary background information for the chapter. In a very few cases, references are made to pertinent sections of other chapters inorder to improve the comprehension of the text and avoid excessive repetition. Except for Chapter 11 which summarizes the transmission line grounding practices of North American utilities, this book can first be divided into five major blocks, namely: • Fundamental Considerations and Practical Design Methodology • Computer Programs • Design Charts • Measurement Techniques andTest Equipment • Concept, Theories andAnalytical Development These blocks overlap in that the same chapter is sometimes referred to in more than one block. Nevertheless, the block structure should satisfy the needs of most readers. For example, the reader who is only interested in design applications need not read Chapters 4 to 8. If an application involves the use of the computer programs, Volume 2 need not be used. The following describes the contents of the five major blocks and explains the interrelations between the blocks and their parts. This should permit the reader todecide where to start in order to best fit his immediate and, perhaps, longer term needs. Fundamental Considerations and Practical Design Methodology Chapters 2, 3 and major portions of Chapter 10 contain practical information regarding transmission line grounding design. These chapters clearly explain how to determine and gather the necessary data required by the four computer programs forming the software package GATE (Grounding Analysis of Transmission Lines) or by the design charts of Volume 2. The reader interested in an immediate application but with insufficient exposure to grounding problems, must first read Chapter 2 on basic concepts and general information. He then should proceed directly to Chapter 3 where a realistic but not overly complex design problem is examined in full detail. The user should participate actively while reading this chapter by personally repeating the computer sessions and/or using the referenced design charts. Once the reader has become familiar with the design methodology, computer programs and/or design charts, he may proceed to Chapter 10 where he will gain additional expertise in using the design tools to analyze more difficult and subtle problems. If necessary, the user should not hesitate to refer to the pertinent sections of the other chapters to enhance his understanding of the issues. Chapters 8 and 9 which deal with measurements and safety aspects respectively, may be of great value to the users of the computer programs and design charts. 1-3 Computer Programs The software package GATL (Grounding Analysis of Transmission line Grounding) is an integral part of this EPRI project. It consists of four independent computer programs; LINPA, RESIST, GTOWER and PATHS. These programs are logically interrelated through the design methodology steps. Instructions for the use of these programs are given in Appendices A to D. Sample computer generated printouts are also provided in these Appendices. Although the programs are very simple to use, Chapters 2, 3 and 10 are a prerequisite for their effective use in the design of grounding installations. Design Charts Volume 2 contains a large number of design charts representing several standard North American transmission line designs. These are intended to provide the reader who is not using the computer programs with an alternate tool to predict transmission line grounding performance with reasonable engineering accuracy. It must be realized that the accuracy of the desi gn charts will decrease as the actual design problem parameters depart from those used in the reference example of the charts. Chapter 3 describes how correction factors can be applied to improve the accuracy in such circumstances. It is clear however, that the use of the design charts are restricted to cases si mil ar to the base examples used to generate the charts. Measurement Techniques and Test Equipment The reader who is interested in the practical aspects of measurement techniques and test equipment should read Chapters 9 and 10 for a complete discussion of this subject. Concepts, Theories and Analytical Development The concepts and theories on which this book and the related software package GATL are based, are described in Chapters 4 through 8. A good knowledge of these chapters is essential for those who wish to improve their analytical expertise in grounding problems or are planning to maintain and enhance the GATL package. It is strongly suggested that the nontheoretical sections of these chapters be read by those who wish to acquire a general understanding of the subject without getting overly involved in the analytical development. Chapter 4 deals with the earth resistivity measurement interpretation techniques. Chapter 3 presents analytical methods to predict the power frequency performance of transmission line grounds while Chapter 6 concentrates on the calculation of the distribution of fault current between structures and overhead ground wires. Chapter 7 is a tutorial on the subject of the response of transmission line structures to lightning strokes. Finally, Chapter 8 describes modern theories relating to the passage of electric current through the body of a living creature and introduces deterministic and probalistic approaches to safety around transmission line structures. 1-4 1.5 CONCLUDING REMARKS Although this book is reasonably complete, there are certain aspects of the problem which require further research work. Lightning performance of transmission line structures is one typical example. Also, the problem of coexistence of transmission lines and pipelines along the same corridor has not been addressed. This exclusion was specified in the Request For Proposal. Because this subject is a natural extension of this project, the analytical methods and computer programs were developed with the possibility in mind of making future enhancements to handle transferred potentials from transmission line structure grounds to neighbouring buried metallic structures such as pipelines. As these lines are written, active research on grounding continues throughout the world. The tremendous capabilties of modern computers have virtually eliminated all of the computational constraints which have severely restricted the progress of power system grounding research for nearly half a century. 1-5 FUNDAMENTAL CONSIDERATIONS 2.1 GENERAL At an early phase of an engineering analysis, it is of utmost importance that all significant parameters influencing the problem be correctly identified. This is particularly true in transmission line grounding studies where a large number of parameters, many of which may vary over wide limits, have a significant influence on the response of a structure during power frequency and lightning fault current events. Consequently, the power system engineer must carefully examine each problem in order to avoid oversimplified or unnecessarily complex analyses. In many cases, simple calculations and sound engineering judgment may be all that is needed to design a transmission line grounding system. In some problems however, the design will require careful measurements and elaborate computer calculations using programs such as those developed for this EPRI project. In both cases however, the engineer must first determine the complexity of the problem in order to decide which methods to use. This initial design step requires some preliminary calculations, using simple formulas, to determine the orders of magnitude of the problems involved. It also requires a knowledge of the fundamental influencing parameters and the capabilities and limitations of the computerized methods likely to be used if accurate calculations are necessary. This chapter contains general information related to transmission line grounding and describes the four computer programs developed for this EPRI project. 2.2 TRANSMISSION LINE DISTURBANCES Ideally, the current in the earth and the current in the neutral conductors of a perfectly balanced three-phase power system are equal to zero during normal operating conditions. In practice however, phase voltages are not perfectly balanced and the transmission line conductor configuration and characteristics are not perfectly symmetrical because of the unequal distances between the phase conductors and earth. Also, various loads on the system are often significantly unbalanced. This means that even during normal conditions, all power systems will have some continuous current in the earth and neutral circuits. This "unbalance current" is generally a few percent of the total transmission line load current. The distribution of this current depends on the particular geometry of the line, number of neutral conductors, and resistance of the ground connections such as substation and transmission line structure grounds. 2-1 In practice, most of the unbalance current enters substation ground grids and returns to the generating source ground system. Very little current normally flows in the neutral wires and transmission line structures because the impedance of the path offered by the structures and neutral wires is typically one or two orders of magnitude higher than a typical station ground resistance. Since the value of a substation's ground resistance is generally a fraction of an ohm, there are usually no safety hazards associated with the unbalance current. The majority of power system disturbances result in the circulation of large currents in the earth. The power system response to such disturbances is largely dependent on the grounding system design and earth characteristics. Two of the transmission line disturbances which occur most frequently are: • Phase to ground faults • Lightning strokes The analysis of a power system subject to a lightning stroke or phase to ground fault at a transmission line structure requires specific data on the overhead network, earth path and nature of the disturbance. Understandably, it is not possible to consider all the parameters which will influence system performance. Rather, those parameters which have significant effects on the electrical response of the power system should be retained. The following variables define the parameters which, according to the recent literature on the subject, play a critical role in the grounding performance of transmission lines. These variables have been grouped into four major categories: • Earth • Power system network • Grounding system • Nature of fault The accuracy of the analyses depend on the accuracy of mathematical models developed from these variables. 2.3 EARTH The development of a mathematical model to represent the electrical properties of earth can be a formidable task because of widely nonuniform characteristics. Fortunately for transmission line grounding purposes, the earth can be reasonably approximated by a two-layered soil structure. This soil structure is characterized by the layer resistivities P i, 92 and the upper layer thickness h. The lower layer is considered infinite. In some cases, the thickness of the upper layer is large enough for the earth model to be considered fairly uniform. The variables Pj, P2 and h are generally determined by interpreting the apparent resistivity values measured using the Wenner (or four probe) array. 2-2 2.3.1 SIMPLIFIED METHODS There are no simple formulas which will define an equivalent two-layer model of a complex earth structure from measurement data. Some useful graphical techniques are explained in Chapters 3 and 4. An equivalent uniform soil is not easy to determine. There is no basis for assuming that a uniform soil model with a resistivity equal to the average measured apparent resistivity will give more accurate results than one using a resistivity value arbitrarily selected from the measured results. The appropriate choice is greatly dependent on the real earth structure and grounding system configuration. Thus, the experience and judgment of the engineer are critical to the development of an equivalent uniform earth model of a complex earth structure. 2.3.2 COMPUTER PROGRAM RESIST The resistivity measurement interpretation process is considerably simplified when program RESIST is used. Detailed information on the use of this program is included in Chapters 3, 4 and 10 and in Appendix B. Because RESIST requires a minimum of data from the user, it can be used by engineers with very little exposure to power system grounding. The input data consist of the probe spacings and apparent resistances (or resistivities) measured using the Wenner method. RESIST automatically selects and calculates all the other data needed to proceed with the final computations. The simplicity and accuracy of program RESIST has been achieved by restricting the applicability of RESIST to transmission line grounding problems. Attempts to apply this program to other types of grounding problems may not lead to optimal solutions. This remark applies equally well to GTOWER and PATHS and, to a lesser extent, to LINPA. For example, RESIST must select initial values for the upper layer thickness and layer resistivities before starting its iterative search algorithm. The initial values retained are appropriate for relatively small grounding systems such as those employed in transmission line structures. Finally, it is paramount to remember that the performance of RESIST depends largely on the number of measurements and extent of the resistivity traverse. A small number of data points or short traverses inevitably lead to a multiplicity of equally probable earth models. Very long traverses unnecessarily emphasize the influence of the deep layers which have very little effect on transmission line grounding performance. In practice, 10 to 20 measurements along a traverse extending 5 to 10 times the transmission line structure base dimension will usually lead to reasonable success. 2.4 POWER SYSTEM NETWORK Throughout this report, it is assumed that the fault or disturbance is at a structure on a transmission line section between two substations. Each substation and the power system network to which it is connected is designated as a terminal. Figure 2.1 illustrates such a typical network. 2-3 PHASE r PHASE q PHASE p TERMINAL Figure 2.1 LOCATION OF FAULT Equivalent Power System Network The terminals can be described by the equivalent circuit of the network which exists at either extremity of the transmission line. The variables which identify the terminal are: V Phase to ground voltage of the terminal power source (if not a power source the voltage is set to zero) Zs Equivalent source impedance of the terminal Zn Equivalent neutral impedance of the terminal Rt Equivalent ground impedance of the terminal The term "equivalent" is used because the substations at the transmission line extremities are always interconnected with other substations. In practice it is usually possible, through suitable circuit reduction, to obtain an equivalent Thevenin circuit which reduces the problem to the basic two terminal circuit shown in Figure 2.1. At times, this reduction is not possible without judicious approximations. Usually, a valid approximation is to replace the network driving the step-down transformer, which feeds the transmission line under consideration, by an infinite source with zero neutral and ground resistances. Because of the distances involved, the effect of this procedure will usually be insignificant at the location of the disturbance. Moreover, if the distance between the origin and extremity terminal is great enough, the terminal busses can be considered as infinite busses (i.e., zero equivalent source impedances) and the equivalent neutral and ground impedances will be equal to those of the terminal transformers feeding the line. The transmission line is assumed to consist of several zones, 1,2,3—m. Each transmission line zone (TLZ) consists of one or more sections. All the structures within a given zone have the same ground impedance value. A transmission line section (TLS) is defined by; the phase conductors, the ground wires between two grounded transmission line structures, and one of these two grounded structures. By convention, the structure which is defined in a given TLS is the structure which is further from an observer as he moves from the point of fault to any terminal. Each TLZ, and consequently, any TLS belonging to this zone is characterized by the following parameters: Zp Self impedance of phase conductor p Zpq Mutual impedance between phase p and phase q Zg Self impedance of ground wire g Zpg Mutual impedance between phase p and ground wire g R Ground impedance of a transmission line structure. The reactive part of this impedance is generally negligible compared to its resistive part. R can be determined from measurements or can be calculated from the structure grounding system configuration and soil characteristics. More will be said about this resistance in Section 2.5. The self and mutual impedances of the overhead conductors of the TLS are very seldom measured directly. Generally, these values are determined from tables and charts or are calculated from simplified formulas or computer programs, usually designated as line parameter programs. 2.4.1 CONDUCTOR IMPEDANCES The self and mutual impedances of parallel cylindrical conductors located above or below the surface of a uniform soil are given by Carson [1] and Pollaczeck [2]. Clarke [3] expresses the self and mutual impedance of two overhead conductors with common earth return as follows: 1=1. + Z P Z « (2-1) ep (2-2) z epg In (2-1) Z; represents the internal impedance of the conductor. This impedance can be determined from conductor data provided by the manufacturer. Zep represents the resistance and reactance of the component of self impedance with earth return external to the conductor p. In (2-2) Zgpg represents the resistance and reactance of the mutual impedance with common earth return between conductors p and g. These impedances are generally expressed in ohms per unit length of conductor, generally a kilometer or a mile. In Equations 2-3, 2-4, 2-6 and 2-7, the impedances are given in ohms/meter. pg Zep and Zepg can be computed accurately from Carson's formulas which involve the summation of tne terms of an infinite series. For frequencies of 60 Hz or less, only a small number of terms is necessary to obtain accurate calculations. Clarke [3J provides simplified formulas and charts for the computation of the external self and mutual impedances at power frequencies. Recently, Deri et al [4] have shown that the Dubanton's [5] simplified analytical expressions for the external earth return self and mutual impedances Zep and Zepg are accurate to within a few percent over a wide range of frequencies. They also analytically relate Dubanton's formulas to those of Carson by showing that the concept of an equivalent complex ground return plane is analytically sound. The Dubanton-Deri formulas are: 2-5 y<3 1 n ep = jw — 2tt 2 (h +D) E J(h epg = ju) — 2ir 1 n (2-3) +h +2D)2 + h2 J__ E_J_________E9 (2-4) V(h -h )2 + h2 f p g pg where w = Z-rf ; f being the frequency in Hz M0 = 47tx10-7 is the magnetic permeability of free space r is the conductor radius hp is the height above the surface of the earth ofconductor hpg is the distance between conductors p and g p D is the complex depth of the equivalent groundreturnplane. If P is the uniform soil resistivity, then D equals: V? D = (2-5) VJWPo The above formulas are convenient for computations using hand calculators and are applicable to low as well as high frequencies. Therefore, they also can be used to calculate the high frequency impedance of counterpoises buried close to the surface (hp ~ 0). If the depth d of the counterpoise is such that the counterpoise can not be considered as lying on the surface or if the counterpoise resistivity and permeability are to be taken into consideration, then the following formulas, described by Wedehpohl [6], are preferable: ! Z. = wy. r j wyj' i V j —coth jO.777^1 j —— r irr 2 jmy ep 1 f /jwy ' r -i ~ ' 1n Y\----- ~ 2 Lip d -1* 2 J 0. 356p: + — •nr /jojy ‘ " ----- 3fpJ where y = 1.781.... (In 7 is Euler's constant) Mi is the magnetic permeability of the conductor Pi is the resistivity of the conductor M is the magnetic permeability of earth P is the resistivity of earth 2-6 (2-6) (2-7) 2.4.2 COMPUTER PROGRAM LINPA Power utilities and consultants usually have access to a line parameter computer program which provides the self and mutual impedances of power conductors arranged in any configuration. Quite often however, the computer results available to the engineer are limited to the equivalent phase sequence parameters. That is, the values obtained after the overhead ground wires have been eliminated by a suitable circuit reduction technique. Because the ground wire impedance and its mutual impedance with the faulted phase conductor play a critical role in the fault current distribution between the ground wire and the faulted structure, it is essential that these impedances be known and represented during the fault calculations. This information is readily obtained from the computer program LINPA which was specifically developed to serve as a convenient and accurate tool for the computation of the impedances required by the fault current distribution computer program PATHS. LINPA is described in Appendix A and is also discussed in Chapters 3 and 10. Computer program LINPA is valid for any configuration of overhead parallel conductors over a uniform earth. The conductors are grouped into different phases, belonging to one or several circuits. If a group of conductors are to be replaced by an equivalent single conductor, the conductors are assigned the same phase number and circuit number. For example, in a transmission line consisting of three phase conductors and two ground wires, the ground wires may be assigned the phase number 4 and LINPA will produce a 4x4 matrix where conductor number 4 will be the equivalent conductor representing the original ground wires. 2.4.3 COMPUTER PROGRAM PATHS PATHS is a short-circuit analysis program which calculates the distribution of current between the overhead ground wires and the faulted structure during transmission line ground faults at that structure. Optionally, PATHS will provide the fault currents in the ground wire spans and in the structures adjacent to the faulted structure. PATHS is described in Appendix D and is discussed in detail in Chapters 3, 6 and 10. Theinput data to PATHSconsist of thevariables illustrated in Figure 2.1. Thesevariables werediscussed earlier. One salient feature ofPATHS isthat it uses the actual 3-phase representation of the circuit, not the zero, positive and negative sequence representation. Thus, when only the sequence impedances are known, these must be converted back to the actual values according to the following equations: 1 Z (2-8) = - (Z0+2ZO s 3 1 Z m = - (Zo-Zi) (2-9) 3 Where Zs and Zm are the actual self and mutual impedances, respectively. Zg and Z\ represent the zero and positive sequence impedances, respectively (assuming equal positive and negative impedances). PATHS provides accurate computation results for single, double and three phase-to-ground faults at any structure along the transmission line being considered. If PATHS is not readily available or if approximate results are desired, then Equations 6-15 and 6-34 of Chapter 6 may be used. These equations provide convenient analytical expressions for the calculation of the fault current in the structure, as a fraction of the total single phase-to-ground fault current. 2-7 2.5 GROUNDING PERFORMANCE OF TRANSMISSION LINE STRUCTURES The distribution of fault current between a structure and the overhead ground wires connected to it is greatly dependent on the ground resistance of the structure. As stated, this resistance R can be either measured or calculated. Knowledge of surface potentials around the faulted structure may also be needed to assess the overall grounding performance of the structure. These earth potentials are usually calculated for local soil characteristics. 2.5.1 SIMPLIFIED FORMULAS The ground resistance of a structure and the earth potentials around it may be calculated using simplified analytical expressions under the assumption of uniform soil resistivity P and simple structure grounding systems such as a hemispherical electrode, a vertical rod and a horizontal cylindrical conductor. These expressions also assume that the current in the electrode discharges into earth uniformly along its length. If these assumptions can be tolerated in a practical design situation, then it is possible to use a variety of approximate formulas available from various publications. Important work on this subject is described in Dwight's paper [7] and Tagg's book [8]. Useful summaries are also provided in IEEE guide 80 [9]. A very useful approach to transmission line grounding design is to initially consider the structure base and associated grounding elements as equivalent to a hemispherical electrode with an appropriate radius r. Under these assumptions, the ground resistance R of the structure is: P R = ----2Trr and the surface potential V at a distance a current I in the earth is: (2-10) x pi V = ----- from the center of the hemisphere injecting (2-11) 2ttx 2.5.2 COMPUTER PROGRAM GTOWER In most practical cases where the earth structure and grounding configuration are complex, simplified formulas based on a uniform soil approach often fail to predict the results, sometimes even within an order of magnitude. The agreement between calculated and measured results are considerably improved when computer programs such as GTOWER are used to determine the transmission line grounding performance. This is well illustrated in Chapter 10 of this report. GTOWER is described in Appendix C and is discussed further in Chapters 3, 5 and 10. GTOWER calculates the performance of a transmission line structure grounding system made up of horizontal and vertical cylindrical conductors located arbitrarily in a uniform or two-layer soil. An optional feature of the program expedites the data entry procedure for the case of a symmetrical grounding arrangement. The program automatically subdivides the conductors into sub-elements to accurately calculate the nonuniform current distribution along the grounding conductors. 2-8 The input data to the program consist of the soil structure data, ground conductor coordinates and earth profile specifications. The output of GTOWER gives the ground resistance and the surface potentials along specified profiles. These profiles can also be plotted if desired. There is one restriction in the program which may not always be identified by the program error-detection routines. Because the potentials are defined only in earth and at the surface of a conductor, requests to calculate potentials just above the center of a ground rod buried at zero depth will cause an arithmetic error. The possibility of this happening will be eliminated if a very small burial depth is assigned to the ground rod. Limits are imposed on the radius of the ground conductors. The radius can not exceed 0.5 m (1.6 ft) for horizontal conductors and 2 m (6.5 ft) for ground rods. There are two good reasons for these limits. Firstly, such a large radius is very uncommon in practice and the error detection capabilities of GTOWER are enhanced by such a limitation. Secondly, the fault current distribution at the surface of a very large radius conductor is not symmetrical because of the proximity effects to other ground conductors. This is contrary to the assumptions on which GTOWER is based. If necessary, this limitation can be alleviated by replacing a large radius conductor with several smaller ones regularly spaced along the surface of the original conductor. In fact this procedure is recommended for a conductor radius above 0.5 m if improved accuracy is desired. 2.6 TYPE OF DISTURBANCE The disturbance is assumed to be localized at a specific structure (structure 0) of the transmission line. This disturbance could be a surge, caused generally by a lightning stroke to the structure, or a power frequency ground fault at the structure. The ground fault could be initiated by various causes such as insulator pollution or a backflash resulting from a lightning stroke. 2.6.1 POWER FREQUENCY GROUND FAULTS Structure faults may be of the single, double or three phase-to-ground type. The overwhelming majority of transmission line power frequency ground faults are single line-to-ground. The type of fault in program PATHS is selected by simply assigning a suitable value to the shunt impedance between the phase conductor and the faulted structure. This impedance may be regarded as the impedance of the arc path developing between the faulted phase and the structure. In most cases, the arc impedance is negligible compared with the other impedances of the circuit. Healthy phases are modelled by simply selecting a shunt impedance high enough to represent an open circuit condition between the phase conductor and the structure. 2.6.2 LIGHTNING STROKES The lightning response of transmission line structures is certainly one of the most difficult subjects in the area of transmission line design. The basic difficulty in calculating lightning performance of structures is due not only to the mathematical effort required to obtain accurate results, but to the uncertainty in the identification of all parameters which have a significant influence on this performance. Moreover, unlike the other subjects considered in this project, there are a number of unresolved controversial issues relating to this matter. 2-9 Unfortunately, recent extensive artificially-triggered lightning test programs have not been capable of resolving the controversies, despite the wealth of new information which is significantly broadening our knowledge on the subject of lightning. Moreover, some of the new evidence may very well be a source of vigorous discussion and additional disagreement as explained in Chapter 7. Chapter 7 contains a brief but complete description of the transmission line structure response to lightning stroke currents. The most widely used analytical approaches are described and commented upon. From this chapter it can be concluded that despite its incompleteness and some deficiencies, the so-called "transient surge impedance" of a structure as given by Wagner [10] remains the most accepted way to represent the lightning response of the structure: Z$ = 60 In [V5 (ct/a)] (2-12) where c is the velocity of light (3x108 m/s) t is the time in seconds a is the radius of the equivalent structure in meters Zs is the surge impedance in ohms Further research work in this area is clearly needed. Accurate methods appear feasible. However, these methods will require the use of advanced and sophisticated techniques only suitable for treatment by digital computers. REFERENCES 1 - J. R. Carson, "Wave Propagation in Overhead Wires With Ground Return", Bell System Technical Journal, No. 5, October 1926. 2 - F. Pollaczek, "Ueber das Feld Finer Unendlich Langen Wechselstrom Durchflossenen Einflachleitung", Elektrische Nachrichten Technik, No. 3, September 1926. 3 - E. Clarke, "Circuit Analysis of AC Power Systems, Volume 1; Symmetrical and Related Components", Book published by General Electric Company, 1961. 4 - A. Deri, G. Tevan, A. Semi yen, A. Castanheira, "The Complex Ground Return Plane; A Simplified Model for Homogeneous and Multi-Layer Earth Return", IEEE Transactions on PAS, Vol. PAS-100, No. 8, August 1981, pp. 3686-3693. 2-10 5 - G. Gary, "Approche Complete de la Propagation multifilaire en haute frequence par Utilisation des Matrices Complexes", EDF Bulletin de la Direction des Etudes et Recherches, Serie B, No. 3/4, 1976, pp. 3-20. 6 - L. M. Wedepohl, D. J. Wilcox, "Transient Analysis of Underground Power Transmission Systems", Proceedings of the IEE, Vol. 120, No. 2, February 1973, pp. 253-260. 7 - H. B. Dwight, "Calculation of Resistances to Ground", Electrical Engineering, Vol. 55, December 1936, pp. 1319-1328. 8 - G. F. Tagg, "Earth Resistances", George Newnes Ltd., London 1964 (book). 9 - Guide IEEE 80 (1976), "Guide for Safety in Substation Grounding", published by IEEE, New York 1976. 10- C. F. Wagner, "A New Approach to the Calculation of the Lightning performance of Transmission Lines. Ill- A Simplified Method: Stroke to Tower", AIEE Transactions on PAS, Part III, Vol. 79, October 1960, pp. 589-603. 2-11 TYPICAL DESIGN METHODOLOGY 3.1 GENERAL There is a wide variety of transmission line grounding problems requiring thorough analysis by power system engineers. Some of these problems can not be fully addressed without concepts and analytical methods not within the original scope of this EPR1 project; transferred potentials (or resistive coupling) to nearby pipelines, structures and buried low voltage circuits, for example. However, most problems can be accurately examined using the analytical methods developed in Chapters 4 to 7. It is recommended that Chapters 4 to 7 be carefully read, although it is not necessary to become familiar with the analytical theories in order to effectively use the computer programs developed from these theories. It is, however, exceedingly important to follow an appropriate methodology when applying the various computer programs to the design of optimum transmission line grounding systems. In this chapter, a typical transmission line grounding design problem is examined in detail using the computer programs developed during this EPRI project. Solutions to the problem based on the design charts of Volume 2 are also provided. Obviously it is impossible to include in this example all the possible variations which emerge in practical problems. However, the design problem of this chapter and the various problems examined in Chapter 10 constitute a good starting point in gaining the necessary experience for effective use of the computer programs and design charts developed during this EPRI project. 3.2 DESCRIPTION OF THE PROBLEM The four-legged lattice tower shown in Figure 3.1 is a typical structure for a planned 345 kV transmission line. One of these towers will be erected near a schoolyard. For this reason, it was decided to design a grounding system which will maintain safe step and touch voltages around this tower. 3-1 AVERAGE HEIGHTS USED TO CALCULATE LINE PARAMETERS 21.5' Figure 3.1 345 kV Four-Legged Lattice Tower 3.2.1 TOWER FOOTING The foundation of each tower leg is of the grillage type shown in Figure 3.2. The grillages provide a good grounding system for the tower. This grounding system can be accurately modelled by equivalent cylindrical elements based on the GMR concept or the approximate "equal cross-section" approach of Chapter 5 (Section 5.2.2). It is assumed that the "equal cross-section" approach leads to an equivalent grillage made of cylindrical elements with a diameter as shown in Figure 3.2(c). 3.2.2 POWER SYSTEM NETWORK The "schoolyard" tower is 3.7 km (2.3 miles) from substation LEFT and 20.9 km (13 miles) from substation RIGHT. The average span between towers is 243.8 m (800 feet). A short-circuit occurs between phase A of circuit No. 1 and the "schoolyard" tower as shown in Figure 3.3. There are 14 towers on the left of the faulted tower and 85 towers on its right. Circuit No. 2 of the transmission line will become operational only 15 years after circuit No. 1 is energized. 3-2 PROFILE LEGEND SOIL SURFACE 5.V VERTICAL ROD i—- a • • JL V CYLINDRICAL CONDUCTOR <p = 0.08' b) Figure 3.2 cj SECTION “DD" EQUIVALENT LEG Grillage Foundations FUTURE Z ~ O.Hl -f j 1.169 fi/mile = 0.09+J0.53 S2/mi Z = 0.808 + a/m] I 2-3 miles 15-3 miles TERMINAL "LEFT" Figure 3.3 TERMINAL "RIGHT" Power System Network 3-3 The phase-to-ground short-circuit levels at both substation busses have been determined, using a conventional short-circuit program, as 8000 MV A for substation LEFT and 3000 MVA for substation RIGHT. These short-circuit levels are converted into approximate equivalent source impedances (as seen from the substation bus) of the network feeding a current into the fault via the substations: (345 kV)2 Z I a1 14.878 Q 8000 MVA (345 kV)2 39.675 ft 3000 MVA It is further assumed that the inductance to resistance ratio of the equivalent source impedance is 15. Therefore: Za = 0.99 + j14.85 ft Zb = 2.64 + j39.59 ft Generally, the substation transformers at each end of the transmission line will contribute the major portion of fault current which will return to the transformer neutral points through the overhead ground wires and substation grounding grids. In many cases however, significant fault current contribution is also provided by other substations. Consequently, the ground impedances Ra and R5 shown in Figure 3.3 are the equivalent ground impedances of all current sources on the left and right of the faulted tower. In practice Ra and R5, typically 0.5 ohm or less, will have a negligible effect on the fault current magnitude and its distribution between the faulted tower and the overhead ground wires. This is why Ra and Rb are assumed equal to 0.01 ohm in this example. The above determination of the equivalent source and ground return impedances is an approximation which leads to accurate results in most cases where the primary source substations are interconnected to a large power network. If a source substation is not connected to other substations, then the source impedance is the equivalent impedance of all power transformers in parallel and the ground return impedance is the substation grid impedance. Finally, it is always possible by use of a circuit reduction algorithm, to reduce the source substations and their interconnections into an equivalent source as shown in Figure 3.3. The equivalent left and right source networks will be designated as "terminal LEFT" and "terminal RIGHT" respectively to distinguish them from the actual substations "LEFT" and "RIGHT". 3.2.3 TRANSMISSION LINE CONDUCTOR DATA Each phase is a bundle of two 954 MCM ACSR conductors as shown in Figure 3.1. The conductors have a radius of 3.04 cm (1.196") and are characterized as follows: Code : Cardinal Stranding : 54x0.1329" aluminum Number of layers : ; 7x0.1783" steel 4 Resistance : 0.0620 ft/Km at 60 Hz and 25 °C (0.0998 ft/mile) GMR : 0.0123 m (0.0404 ft) Reactance at 1 ft radius : 0.242 ft/Km at 60 Hz (0.389 ft/mile) 3-4 There are two 19 No. 10 alumoweld overhead ground wires which have the following characteristics: Conductor diameter : 1.29 cm (0.509") Resistance : 0.893 fl/Km at 60 Hz and 25 °C (1.437 ft/mile) GMR : 0.000841 m (0.00276 ft) Reactance at 1 ft radius : 0.444 ft/Km at 60 Hz (0.715 ft/mile) The average height above ground of the phase conductors and ground wires is assumed equal to the height at the tower (see Figure 3.1) less half the average conductor sag which is about 22 ft. Finally the average soil resistivity along the transmission line right of way is estimated at 150 ft-m. 3.2.4 APPARENT SOIL RESISTIVITY The apparent soil resistivity around the "schoolyard" tower was measured using the Wenner method. The results of the measurements are given in Table 3.1. PROBE SPACING (m) Table 3.1 APPARENT RESISTIVITY (ft-m) 2.5 320 5.0 245 7.5 182 10.0 162 12.5 168 15.0 152 Apparent Soil Resistivity 3.3 DESIGN METHODOLOGY The main objective is to design a tower grounding system which limits the touch and step voltages around the tower to safe values. This can be accomplished by following the steps shown in the flowchart of Figure 3.4. In this flowchart, the bracketed names on the right-hand side refer to the computer programs which can be used to perform the necessary calculations (see Appendices A to D for a description of these programs), while the left-hand bracketed descriptions refer to design charts contained in Volume 2 of this report or to simple equations described in this volume. The flowchart describes the three principal methods which can be used to mitigate step and touch voltages. The subjects of safety and mitigation measures are discussed in Chapter 8. 3-5 3.3.1 TWO-LAYER EARTH MODEL Table 3.1 clearly shows that the soil around the tower is not uniform. Therefore, an equivalent two-layer earth model must be determined to improve the accuracy of the tower grounding performance calculation. The resistivity values of Table 3.1 were used as input data to computer program RESIST as shown in Appendix B. This appendix provides instructions for preparing the input data file (for batch mode processing) and the input data dialogue when the interactive mode is selected to run the program. The computer printout of this appendix shows that the equivalent earth structure determined by RESIST corresponds to a 2.56 m (8.4 feet) thick first layer with a resistivity of 383 fi-m underlain by a second layer with a lower resistivity value of 147.7 Q-m. An almost identical two-layer earth model can be determined using the logarithmic curve matching technique described in Chapter 4. The field resistivity curve is plotted on a transparent logarithmic graph and is then compared directly to the set of theoretical Master curves provided in Volume 2 (Figure 1.1). The comparison process, illustrated in Figure 3.5, consists of obtaining a satisfactory match between the measured and theoretical curves through a series of appropriate horizontal and vertical translations of the transparent graph sheet. When this is accomplished, the thickness of the upper layer is the value of the abscissa on the vertical line passing through the Master chart abscissa corresponding to a/h = 1. From Figure 3.5 it is found that the upper layer thickness is approximately 2.5 m (8.2 feet). Similarly, the upper soil resistivity is the value of the ordinate on the horizontal line passing through the Master chart ordinate corresponding to Pa/P| = 1. Figure 3.5 shows that the upper soil resistivity is about 390 fi-m. Finally, this figure indicates that the measured apparent resistivity curve corresponds to a reflection factor in the range of -0.4 to -0.45. This leads to a lower resistivity value between 148 and 167 Q-m. 3.3.2 TRANSMISSION LINE PARAMETERS The self and mutual impedances of the transmission line phase conductors and ground wires are important parameters for the calculation of the fault current and its distribution between the faulted tower and the overhead ground wires. In particular, the self impedance of the ground wire and its mutual impedance with the faulted phase conductor have a significant influence on the portion of fault current which flows down the faulted tower. These impedances are not always readily available to the design engineer and must be calculated using transmission line parameter computer programs, charts or approximate formulas such as those given in Chapter 2. Computer program LIN PA is a line parameter program specifically developed to calculate the line parameters required by the computer program PATHS, which determines the fault current and its distribution between towers and ground wires. Appendix A provides the instructions to use LINPA in a batch or interactive mode. It also gives the complete printout of the LINPA run based on the design problem data described in Section 3.2.3. The results of this run are summarized in Table 3.2 which gives the symmetrical impedance matrix (excluding circuit No. 2) after the conductors of a bundle and the ground wires have been reduced to an equivalent conductor. The bundle and ground wire reductions are essential since the computer program PATHS uses the self and the mutual impedances of the equivalent phase conductors and equivalent ground wire. 3-6 APPARENT SOIL MEASUREMENTS VOLUME 2 PART 1 DETERMINE EQUIVALENT TWO-LAYER EARTH MODEL RESIST TOWER CONFIGURATION AND CONDUCTOR CHARACTERISTICS VOLUME 1 CHAPTER 2 DETERMINE TRANSMISSION LINPA PARAMETERS CONFIGURATION OF TOWER GROUNDING VOLUME 2 PARTS 2,3, AND k DETERMINE TOWER GROUNDING PERFORMANCE GTOWER CURRENT IN TOWER USE LOW IMPEDANCE OVERHEAD GROUND WIRES iVOLUME 2) ;PART 3 / VOLUME 1 1 ^CHAPTER 6' DESIGN COMPLETED DETERMINE FAULT CURRENT IN TOWER AND IN OVERHEAD GROUND WIRES OPTIMIZE GROUNDING SYSTEM CONFIGURATION (GROUND POTENTIAL CONTROL RINGS, COUNTERPOISES, ETC.) PATHS IS TOWER GROUNDING SYSTEM \ SAFE ? / LAYER ON SURFACE OF SOIL SELECT MITIGATION MEASURES Figure 3.4 Design Methodology As can be deduced in Appendix A, conductors to be reduced to an equivalent conductor by program LINPA are assigned the same phase number (phase No. 4 for the two ground wires of this design problem) to indicate that these conductors are at the same potential with respect to remote ground. Therefore, the word "phase" in LINPA has a broad meaning. Note that the transmission line impedance values shown in Figure 3.3 and used as input data for the program PATHS (Appendix D) are average values determined from Table 3.2. 3-7 r r I r (ooo- l f t it 390 ft-m b f i b b l i r i i F [ v \ F i f Ff i J2.5 m Figure 3.5 ijml Logarithmic Curve Matching Technique EQUIVALENT CONDUCTOR SELF AND MUTUAL IMPEDANCE A1 * 0 1 2 + jl.168 B1 0.092 + jO.611 0.141 + jl.169 Cl 0.091 + j0.527 0.091 + j0.599 0.141 + jl.169 0.091 + j0.485 0.090 + j0.521 0.090 + j0.602 * A1 = Phase A of Circuit No. 1 GW = Ground Wire Table 3.2 GW* Cl B1 A1 * GW* . A (IN OHMS/MILE) Computed Impedance Matrix 3-8 0.808 + 1.234 3.3.3 TOWER GROUNDING PERFORMANCE The response of a grounding system conducting power frequency current into earth is proportional to the current magnitude unless soil breakdown occurs due to very high current densities in the ground conductors. This situation very seldom occurs at transmission line structures subject to power frequency faults. It is therefore possible to analyze the performance of a grounding system assuming a 1 per unit (p.u.) current and applying appropriate proportionality factors to determine the response of the grounding system to actual or predicted currents. The grounding performance of the tower footings illustrated in Figure 3.2 was determined using the computer program GTOWER described in Appendix C. The same results were derived from the charts provided in Volume 2. In each case the actual current in the tower was not yet known. A value of 1 kA was assumed in the GTOWER run. The design charts are normalized (in %) with respect to the tower current, upper soil resistivity and transmission line configuration and dimensions. The computation results of GTOWER provide the tower ground resistance and earth surface potentials along the profile shown in Figure 3.2 (Appendix C shows the complete printout of a GTOWER run relating to a grounding system consisting of the tower footings and a square ground potential control loop). The worst touch and step voltages occur 0.9 m (3 feet) from the tower leg as illustrated in Figure 3.6. The computed results are as follows: Tower ground resistance : 31.67 ohms Touch voltage Vq (also step voltage VS]J) : 69.1% Touch voltage Vt2 (also step voltage Vs2) : 67.2% The step and touch voltages are expressed in percent of the tower potential rise which is the product of the tower ground resistance and the fault current injected into earth by the tower. Step and touch voltages at other locations around the tower can be determined from GTOWER results if appropriate potential profiles are specified. 3' -- 3' TOUCH VOLTAGE Figure 3.6 STEP VOLTAGE Worst Touch and Step Voltages 3-9 Similar results can also be determined with reasonable accuracy from the charts of Volume 2 (Part 2) because the tower footings correspond to one of the two types provided in Volume 2. It should be noted, however, that although the general shape is the same, the various dimensions between the actual and example towers are not in the same proportion. More will be said about this later. For the moment however, it is assumed that all dimensions of the actual tower footings are proportional to those of the reference tower used to develop the charts of Volume 2. This dimension scale factor Xjj, is defined as the ratio of the tower base dimension (51.8 feet) in the example, to the tower base dimension of the actual tower (28 feet). Therefore is equal to 1.85. A close examination of the equations of Chapter 5 reveals that if all dimensions, including the thickness of the upper soil layer, are scaled up or down by a factor X£ , the response of the grounding system remains inversely proportional to the scale factor Xjj, . It also shows that this response is proportional to the upper layer resistivity P i, provided that the reflection factor is the same. These observations are summarized by the following expression: p actual = p model _ X (1-1) u P where P is the ground resistance or the earth potential (i.e.,response of the grounding system). X£ is the dimension scale factor as defined earlier. Xp is the resistivity scale factor and is the ratio of upper layer soil resistivities in the example and the actual case. The upper layer soil resistivity of the example used in Volume 2 is 100 fl-m. Thus the resistivity scale factor for the design problem is 0.261, reducing the above expression to: P actual =7 ?26 P model Ground Resistance The ground resistance for the example (base) tower is obtained from the chart of Figure 2.3 in Volume 2. The upper layer of the actual soil is 2.56 m deep. This corresponds to a 4.7 m (2.56x1.85) thick upper soil layer in the example. From Figure 2.3, the ground resistance of the tower footings without a potential control conductor in a two-layer soil with a reflection factor of -0.44 (as computed by RESIST) is about 3.4 ohms. Hence, the actual tower resistance is: R aCT.Ua I = 7 • 085x3 . k = 24.1 ohms The resistance value has been derived assuming that all dimensions of the actual tower are proportional to those of the base tower. This is not the case, as can be concluded from an examination of Figure 3.2 and Figure 2.1 of Volume 2. Thus, a correction factor should be applied to account for the nonuniform scaling between the actual and base towers. A logical way to determine this correction factor is to calculate an average dimension scale factor based on the dimensions of the fundamental elements of the grounding system. In the case of this problem the following average scale factor is selected: , A£a = 1 / Li mode 1 3 \ L i actual -I Wi , L? model L3 model Lz actual L3 actua1 ----------- + w2 —------------ + w3 —r----------- 1 / 5L! t J0JL t ju8 \ 3 \ 28 5A 1.6 / 3-10 )' 2 26 (3-2) where Lj_, L2 and L3 are the dimensions shown in Figure 3.7 and wj, W2, W3 are weighting coefficients which describe the relative influence of a ground element and its effect on the variable being calculated. For tower ground impedance, it is considered that the three dimensions selected have equal influence w^ = W2 = W3 = 1. Figure 3.7 Fundamental Tower Grounding System Dimensions The correction factor is therefore 2.26/1.85 = 1.22. The tower ground resistance thus becomes: "actual " 21,.1x1.22 = 29.; ohms The derivation of correction factors to account for minor differences between actual and base tower ground systems will generally be based on experience and engineering judgement. For example, the same weighting coefficient should be applied to all grounding elements if the tower ground resistance is being calculated. If however, a surface potential is being investigated, then it may be more appropriate to apply different weighting coefficients depending on the distance between an element and the point where the potential is calculated. It is important to be cautious when correction factors are used to account for significant differences in scale. Above a value of 1.5, correction factors are suspect, suggesting that the charts of Volume 2 can not be used for such cases. Touch and Step_Voltages The touch (and step) voltages Vq and can similarly be determined from the normalized potential profiles of Figure 3.8 which is also Figure 2.8 of Volume 2. A distance of 0.9 m (3 feet) from the leg of the actual tower becomes 0.9x1.85 = 1.7 m (5.5 feet) from the base tower leg. This is because the earth surface potentials, expressed in percent of the tower potential rise, are the same for two geometrically equivalent points 3-11 of the actual and base towers. Thus, from Figure 3.8 the following values are determined: vt1 = Vs1 = 100^ -35% = 65% Vt2 = VS2 = 100^ "38^ = 62^ Because the above voltages represent the difference in surface potentials at points very close to the tower leg, the influence of differences between the footing grillage of the model and actual towers is negligible. Consequently, the correction factor to be applied on the above voltages is: 1 ) 51.8 + 10.5 2 1 28 5.4 (1.85) = 1.025 735 kV (QH) TRANSMISS ON LINE STRUCTURE SOIL STRUCTURE: Un i form h = 5 U Two- layerLx.1 m POTENTIAL CONTROL RING: YES □ POTENTIAL PROFILE No. : NO SI (Leg) POTENTIAL i r 0 5 , i 1 10 TOWER LEG ■ r ' 15 20 25 Distance from Origin of Profile Figure 3.8 Earth Surface Potential Profiles (in meters) 3.3.4 FAULT CURRENT DISTRIBUTION The magnitude of the fault current and its distribution between the faulted tower and the ground wires was determined by computer program PATHS described in Appendix D. The input data file, the input data dialogue and the computer program results related to this design problem are presented in this appendix. The input data used by PATHS have been described previously and is summarized in Figure 3.3. Note that the tower ground resistance ^e> has been determined by GTOWER and is equal to 31.67 ohms. Note also that the self impedance of the phase conductors as used by PATHS is the average value of phases A, B and C. The computation with only 377 A the soil directly volts. The worst potential rise or results from PATHS show that the fault current magnitude is 15,321 A flowing down the tower. Therefore, only 2.46% of the fault current enters at the faulted tower and the tower potential rise is 377x31.67 = 11,940 touch (and step) voltage previously calculated was 69.1% of the tower 0.691x11,940 = 8,250 volts. A similar result can be obtained using the design charts of Volume 2, Part 5. However, these charts provide the tower current in percent of the total fault current which, if not known, must be calculated by conventional short circuit programs or by other means such as the one which follows. If the coupling between the phase conductors and the ground wires is neglected and if the influence of the nonfaulted phases are also neglected, a simple circuit reduction of the network of Figure 3.3 leads to the equivalent circuit of Figure 3.9. The total fault current If can easily be calculated from this circuit. A further simplification can be introduced by considering that the impedance of the current return path is zero (Re in parallel with Zi and Z2 is negligible). This last assumption is reasonably valid for this problem and leads to a total fault current of about 14,800 A. 0.99+j14.8Q 0.33+j2.7^ +j 18.0^ 2.1 7 3^5 kV /3 345 /3 0 2.64+j 39.6f2 ft Figure 3.9 0 n Simplified Circuit The current It in the tower depends on the ground resistance of the adjacent towers, the impedance of the overhead ground wires and their mutual impedance with the phase conductor. The self and mutual impedances of the ground wire match reasonably well those of the 500 kV transmission line example of Volume 2 (Part 5). As it can be concluded from Chapter 6, the voltage of the transmission line is of no consequence when the current in the tower is expressed in percent of the fault current. Of primary importance are the ground wire self and mutual impedances and the ground resistances of the faulted tower and the immediately adjacent towers. The average span distance and the average ground resistance of the remote towers do influence the results but only slightly. 3-13 The charts which are most applicable to the case being analyzed are Figures 5.41, 5.43 and 5.44 of Volume 2 Part 5. These charts are directly applicable to a 20 mile long transmission line. From Figure 5.41, it can be determined that for an average tower resistance of 20 ohms, the current Ij- in tower 14 would be about 3.6% of the total fault current for a fault at that tower. This result assumes that the faulted tower resistance Re is equal to the average resistance R. Since this is not the case, a correction factor is needed. This factor can be determined from Figures 5.43 and 5.44 of Volume 2. A close examination of these two figures indicates that for R < Re, the following relationship is applicable: It(R<Re) = — It(R=Re) (3-3) R e Thus, the correction factor is, 20/31.67 = 0.63, and the adjusted tower current is, 3.6x0.63 = 2.3%, a value which compares favorably with the results of the PATHS run. Once more, the importance of appropriate correction factors is clearly illustrated. These factors are needed whenever a design problem is different from the reference problem of Volume 2. 3.3.5 SAFETY EVALUATION At this point, the hazard level around the tower can be examined. There are several approaches which can be followed to conduct a safety evaluation. These are explained in Chapter 8. In the case of this problem, Dalziel's well known formula is used to illustrate the design concept: 0.116 (3-4) A where ij-, is the maximum permissible body current in amperes and t is the fault duration in seconds. If Vt designates a touch voltage and Vs represents a step voltage, then safety is achieved when: V < s 116 + °-»2 a = u *f . s (3-5) „.6) /t The above expressions assume that the resistance of the human body is 1000 ohms (see Chapter 8). Rf is the resistance of one foot of a person standing on a thin semi-insulating layer covering the soil surface. From Equation 8-8 of Chapter 8, the value of Rf for a 200 cm2 foot (31 in2) is: Rf = 50eps (3-7) where e is the thickness of the semi-insulating layer (in m) and Ps the resistivity of this layer (in ft-m). Table 3.3 shows tolerable step and touch potentials for several values of the fault duration t, resistivity ps, and surface layer thickness e. No TIME (s) TOUCH VOLTAGE ut (VOLTS) STEP VOLTAGE us (VOLTS) Table 3.3 GRAVEL (ps = 3000 -m) SURFACE e=5cm LAYER (p ASPHALT > 10,000 Q-m) e=10cm e=15cm e=20cm e=15cm e=20cm 0.1 367 1742 3118 4494 5869 4952 9537 14123 18708 0.2 259 1232 2205 3177 4150 3502 6744 9986 13228 0.3 212 1006 1800 2594 3388 2859 5506 8154 10801 0.5 164 779 1394 2010 2624 2215 4265 6316 8366 0.1 367 5869 11371 16874 22376 1 8708 37049 55390 73732 0.2 259 4150 8041 11932 15822 1 3228 26198 39167 52136 0.3 212 3389 6565 9742 12919 10801 21390 31980 42569 0.5 164 2625 5085 7546 10007 8366 18525 24771 32974 e=5cm e=10cm Permissible Contact Voltages The zone within the heavy line of Table 3.3 identifies the cases where safety criteria are not satisfied based on the 8,250 volt worst case touch and step voltages determined from the computer runs related to this problem. It is clear that safety is achieved only through the use of a thick high resistivity surface layer such as asphalt. It is also clear that permissible step voltages are significantly higher than the permissible touch voltages when there is a semi-insulating surface layer. 3.4 POTENTIAL CONTROL CONDUCTORS A touch or step voltage of 8000 volts may not be tolerable in some cases. Potential control conductors could then be used to reduce the voltage to a more acceptable value. In the following paragraphs, the mitigating effects of the square potential control loops of Figure 3.10 are analyzed as these loops are installed in a three step sequence, illustrated in Figure 3.10. The methodology of the analysis is a repetition of what has been described in Sections 3.3.3 and 3.3.4. Note that the design charts of Volume 2 (Part 2) can only be applied to step number 1, i.e., when the potential control loop shown as a solid line in Figure 3.10 is installed. For the next two steps, the computer program GTOWER must be used to analyze the performance of the tower grounding system. 3.4.1 INNER POTENTIAL CONTROL LOOP; STEP 1 Appendix C gives the input data instructions, input data file and computer printout of a GTOWER run based on a tower grounding system consisting of the footings and an inner potential control loop. The computation results show that the tower ground resistance is reduced to 12.17 ohms and the touch and step voltages defined in Figure 3.6 are: vtl = vsl = icm -74.4% = 25.6? Vt2 = VS2 = ]00Z "63.0% = 37.0% Comparable results are obtained from the design charts of Volume 2 by following the same technique used in Section 3.3.3. 3-15 OUTER POTENTIAL CONTROL LOOP (STEP 3) INNER POTENTIAL CONTROL LOOP (STEP 1 LEG POTENTIAL CONTROL LOOP (STEP 2) TOWER LEG CENTER OF TOWER Figure 3.10 Ground Potential Control Loops From Figure 2.3 of Volume 2, the tower ground resistance of the example for a 4.7 m thick upper layer and a reflection factor of -0.44 is about 1.23 ohms. The actual resistance is therefore: '383 x(l.85)x(1.25) =7-085x1.25 = 8.85 ohms 100 This resistance should be further multiplied by the following correction factor: 1 HA + J£i5 + 4-8 +_95 4 1 28 5.4 1.6 37 (1.85) = 1.26 In the above expression 95/37 represents the ratio of the inner potential loop dimensions of the base and actual towers. The adjusted ground resistance of the tower is therefore 8.85x1.26 = 11.2 ohms which is in reasonable agreement with the value computed by GTOWER. Note that it is also possible to account for the influence of the potential control loop dimension by using Figure 2.4 of Volume 2. The results of program PATHS for a fault at the tower with only an inner potential control loop show that the total fault current has remained practically constant at about 15,346 A. The current in the tower has increased to 938 A, however. This corresponds to a potential rise of 938x12.17 = 11,415 volts, i.e., slightly less than the case with no inner loop. The worst touch and step voltage is 37%x 11,415 = 4224 volts, about half the value computed without the loop. If the fault clearing time is known, Table 3.3 can be used to select the thickness and material type of the semi-insulating surface layer which should be used to insure safety around the tower. A similar worst case touch voltage value can be determined from Figures 2.24, 2.17, 5.42, 5.43 and 5.44 of Volume 2. The methodology is identical to that followed in Sections 3.3.3 and 3.3.4. 3-16 3.4.2 LEG CONTROL LOOP; STEP 2 As shown in the previous section, a person touching the leg of the tower while standing underneath the tower, 3 feet from the leg will be subject to a touch voltage of 37% of the tower potential rise. The touch voltage will be only 25% if the person is on the outside (see Figure 3.6). The addition of L-shaped conductors at each leg improves this situation as can be concluded from the following computer results: GTOWER Resistance of tower : 11.03 ohms Touch and step voltages (in % of tower potential rise) Vti = Vsi = 20% Vt2 = Vs2 = 18% PATHS Total fault current : 15,350 A Current in tower s 1,026 A Tower potential rise : 11,324 volts The worst touch voltage has been reduced to 20%xll,324 = 2,265 volts. 3.4.3 OUTSIDE POTENTIAL LOOP; STEP 3 The addition of the outer ground potential loop leads to the following results as computed by GTOWER and PATHS. GTOWER Resistance of tower : 7.34 ohms Touch and step voltages (in % of tower potential rise) Vtl = Vsl = M.2% Vt2 = VS2 = 14.6% PATHS Total fault current : 15,369 A Current in tower : 1,482 A Tower potential rise : 10,876 volts The worst touch voltage is now equal to 14.6%x 10,876 = 1,588 volts. This value is significantly less than the unmitigated value of 8,250 volts computed in Section 3.3.4. 3-17 3.5 OTHER CASES It is obvious that despite the large number of charts available in Volume 2, their use is restricted to cases which belong to the same category as the referrence example used to develop the charts. Deviations from the reference example can be handled by use of a variety of appropriate correction factors, but the judicious choice of a correction factor is essentially a matter of experience. In contrast, the computer programs offer the possibility of analyzing a large variety of configurations without the need for correction factors. GTOWER can be used to explore new potential control loop designs and PATHS, which is valid for single or double circuit transmission lines, can be used to analyze ground faults on any phase; single, double, or three-phase to ground and at any tower. 3-18 CHAPTER 4 EARTH RESISTIVITY 4.1 GENERAL The ability of a group of buried metallic conductors such as transmission structure grounds and counterpoises, to conduct current into the soil is significantly dependent on the resistivity of the soil. Other factors, such as the impedance of power cables and overhead conductors, are influenced by earth resistivity. Earth resistivity can even affect the susceptibility of a particular location to lightning strikes [1,2]. Thus, a knowledge of soil resistivities is of fundamental importance in accurate prediction of transmission line performance during ground fault or lightning conditions. Unfortunately, it is not practical to determine soil resistivity everywhere along the route of a transmission line. There are certain circumstances, identified in Chapter 8, which require that accurate knowledge of soil structure be determined at specific sites. At most other sites, accurate values are not necessary and an order of magnitude estimate is sufficient. Fortunately, there are usually indirect sources of information from which it is possible to secure a qualitative knowledge of the soil structure. This chapter is directed at the measurement and interpretation techniques most generally used to determine soil structure and resistivity. The complexities of most classical methods used by geologists to measure and interpret earth resistivity are beyond the scope of this report. The objectives of a power system design engineer are quite different from those of a geologist, however. Electrical exploration of the earth for geophysical purposes is directed at a determination of soil structure and composition at great depths. In contrast, the electrical engineer needs to know the soil structure at relatively shallow depths and over comparatively short distances, because the performance of grounding structures is influenced mainly by the characteristics of the soil surrounding the grounding electrode. The only information required for grounding purposes is the actual soil structure or an equivalent model resulting in similar performance. This is fortunate because the actual earth can often be modelled as an equivalent two-layer soil structure [3-5], or in some cases, a three or four-layered structure. Because of the complexity of the mathematics required to handle models with more than two-layers, these cases will not be analysed in detail. A two-layer earth model is usually more suitable than an approach based on uniform soil models because the only cases where uniform soil models can lead to accurate answers are those cases where measurements confirm that the earth is in fact fairly uniform. Unlike most engineering problems, interpretation of earth resistivity measurements is an "inverse" problem; i.e., from the electrical response to impressed current at specific locations on the earth surface, the electrical properties of the conducting media (earth) are to be determined. In contrast, conventional electrostatic problems determine the electrical response or the excitation current sources, based on the known properties of the conducting material. These are known as the Laplace and Dirichlet problems. Obviously, the inverse problem, where the physical constants of the material are unknown, presents more difficulties than those problems where the physical constants of the material are known functions of position. 4-1 Geophysical prospecting is an example of the inverse problem, because of the enormous variety of soil structures and characteristics. Moreover, the number of parameters required to represent a model of the earth structure is usually so great that it is difficult to choose initial values to these parameters and have a computer algorithm converge to an acceptable solution within a practical time frame. Consequently, the selection of initial values becomes a fundamental task in the interpretation process. Perhaps this is why geoeiectrical sounding is still considered an art accessible only to experienced geophysicists despite the intricate mathematical theories developed to support this engineering science. Success or failure in this important initial assessment is generally dependent on the experience of the engineer and the knowledge of earth electrical properties available to the engineer responsible for the interpretation of the measurements. There is one further problem with the inverse solution of resistivity measurements. It is not always possible to obtain a unique solution to a data interpretation problem. Because of inaccuracies in the measurements {usually ±5% with classical geoelectric instruments), several models of earth structure can be found to give satisfactory agreement with the measured results. These models will usually differ in the characteristics of the deep soil layers. The above discussions are not presented to discourage the power system engineer from performing a scientific interpretation of resistivity measurements, but rather to make him aware that this task requires careful preparation, investigation, and engineering judgment. The difficulties mentioned previously, while imposing a considerable challenge to the geologist, have significantly smaller impact on the electrical engineer. Firstly, the existence of multiple solutions to the substratum structure is of little consequence in determining the response of ground electrodes, particularly those of the transmission line structure grounds. Secondly, a two-layer earth model is generally sufficient for modelling transmission line structure grounding systems. Finally, there are numerous charts, algorithms, and simple engineering visual estimation techniques which can be used to determine an equivalent two-layer earth model with reasonable accuracy. The following paragraphs are intended to provide the electrical engineer with the basic knowledge and engineering tools required to properly interpret earth resistivity measurements. Initially, the basic electrical properties of earth materials will be described and discussed, since these are fundamental to the interpretation process. Next, measurement methods will be introduced and the classical Wenner method will be examined in detail. The fundamental equations governing the response of the current- carrying electrodes used in the Wenner configuration will then be derived. These equations are essential to the derivation of an accurate earth model. Several interpretation methods will be described using actual measurement results from various sites. Finally, measuring equipment will be described together with problems likely to be encountered in the field. Solutions to the problems will be proposed. 4.2 ELECTRICAL PROPERTIES OF EARTH The conduction of electricity in earth may take place by electronic or ionic current flow. The resistance offered by a material to current flow is expressed in terms of resistivity. This resistivity p is defined by the mathematical expression of Ohm's law: E = pJ (^-l) where E is the electric field strength expressed in V/m and J is the current density in A/m2. 4-2 The unit for resistivity is the ohm-meter. The two opposing faces of a 1 meter cube of a material will exhibit an ohmic resistance corresponding to the value of resistivity of the material. That is a cube of 10 ohm-meter material will have a 10 ohm resistance in this configuration. 4.2.1 CONDUCTION BY ELECTRONIC PROCESS Electronic conduction is characterized by the movement of free electrons in a metallic or semiconductor material when a dc field is impressed on it. Although electrons in semiconductors are only mobile over relatively short distances, no other atomic particles participate in the conduction process. Gold and copper are typical metallic conductors. Another conductor is carbon, which occurs commonly in the form of graphite. The resistivity values of some common refined and unrefined metals are listed in Table 4.1. Because of differences in atomic structures, the resistivities of semiconductors are higher than those of metals. A clearer dishinguishing feature dividing semiconductors and metallic conductors is their temperature behavior. With semiconductors, resistivity decreases with temperature rise at least within a certain temperature range, while the opposite is true with metals. Also, the resistivity of semiconductors varies over a wide range of values. Some typical semiconductors and their average resistivity values are also given in Table 4.1. Metals and semiconductors occur infrequently, when compared to the amounts of other materials in earth. Nonetheless, their occurence in small amounts, particularly in a pure metallic state, may significantly influence the results of apparent resistivity profiles. However, it would be unusual to discover disturbing effects due to native metals during surveys for power engineering applications. This is because the extent of the resistivity traverses is generally too short to penetrate into the deep soil. In the majority of cases, if the presence of a metallic mass is a problem, a man-made metallic structure buried in the vicinity (pipes, etc.) will be the cause. 4.2.2 ELECTROLYTIC CONDUCTION All rocks at the surface of earth are at least slightly porous, and the degree of this porosity helps determine the electrical properties of the rock. Igneous and dense carbonate rocks are not very porous, generally less than 1% of total volume. Other highly porous rocks such as poorly compacted mudstones may be up to 50% porous. The voids are usually completely or partially filled with water having some concentration of salt, and thus can be highly conductive. Depending on the water content, salt concentration, type of metal and its concentration in the rock, the electrolytic conduction process in rocks will be variable. Typically, for surface rocks, electrolytic conduction will be the predominant factor. However, in completely frozen rocks or in deep rocks subject to high overburden pressures, electronic conduction becomes significantly more important than ionic electrical conduction. Also, the presence of magnetite, graphite or pyrite, even in quantities of only a few percent, may reduce the resistivity of a rock to less than 1 ohm-meter if suitably distributed throughout the rock volume. As a general rule however, the resistivity of materials near the earth's surface is dependent on water content and the nature of the salts dissolved in the water. That is why it is not possible to assign a single value for the resistivity of the rock and soil materials shown in Table 4.1. 4-3 TYPE RESISTIVITY Q-m MATERIAL REFINED METALS TYPE SI 1ver 1.5x10'8 Chemiccally clean water 25x101' 1.6xl(f8 Distilled water 5000 Gold 2.0x10~6 Rain water 100 to 1000 Surface water (1ake,rivers) 100 to 500 Aluminum 2.5x10-8 WATER 8.5x10’8 Sea water 0.1 to 1 1 ron 9-xlo"8 Loams, garden soils 5 to 50 Lead 19.x10‘8 Clay, cha1k 10 to 70 Bismuth 100.xio"8 Clay, sand 5 gravel mixtures L0 to 250 Copper 1.2xl0'8 to 30x10~8 Peat, marsh soil & cult!vated sol 1 50 to 250 Diabase, shale, limestone & sandstone 100 to 500 Sand, Cambrian limestone & sandstone lx103 to 3x103 Graphite depending on direction of current flow with respect to cleavage CONDUCTING MINERALS *Note: Table 4.1 TYPICAL VALUES 28x1o'8 to o 5.5xto“8 LIthium o Z i nc (FeS ) Pyrite -6 -8 0.3x10 to 50x10 1.2x10"3to 600x10'S (MoS ) Molybdenite 0.08 to 7.5 Igneous, rocks granite 3x105 to 4x107 (CU 0} Cuprite 10 to 50 Wet concrete 50 to 100 (Fe 0 ) Magnetite 52x!0*6 Dry concrete 2x103 to 1xIO4 (CUO) Malaconite 6000 (PBS) Galena SEMI- Ground water, well and spring water o k. 3xl0~8 SOILS MINERALS RESISTIVITY Q-m Copper Sod 1um NATIVE MATERIAL Moraine, quaternary surface Ix103 to lx104 coarse sand 6 gravel The larger range in values is due to the variations in the composition of the materials. Typical Resistivity Values However, this is not of immediate interest to the power system engineer. The important point is that the absence of water in rocks results in very high resistivities. The rock resistivity will continue decreasing with increasing volume of water, approaching that of the water in the rock for higher porous rock conditions. Temperature variations also affect the resistivity of surface rocks. If the temperature changes so that it becomes high or low enough to evaporate or freeze the water contained in the voids, the resistivity may be expected to increase manyfold. If no evaporation occurs, an increase in the temperature will generally result in a slight decrease in the rock resistivity, due to the increased mobility of the ions in the water solution. Figures 4.1 and 4.2 show the effect of temperature on the resistivity of soils and rock. It should be noted that the resistivity variations of rocks with temperature will not follow those of pure water, since the presence of dissolved salts and the confining effect of small pores tend to lower the freezing point of the water in the rock. For accuracies within an order of magnitude, Hummel's empirical formula, is useful to estimate the resistivity of most soils materials which conduct electricity principally by the electrolytic process: where Ps is the soil resistivity in volume of water contained in soil. -m, Pw is the water resistivity and C is the relative BIOTITE GRANITE .... SAND & GRAVEL SILT TEMPERATURE Figure 4.1 (°C Soil Resistivity Versus Temperature (Redrawn from [6]) 4.2.3 ELECTRICAL CONDUCTION IN EARTH Among the electrical characteristics of the earth, only the relative permeability ^r may be considered as constant and equal to unity. The relative permittivity er and the resistivity p vary within wide limits, although the variations of p (typically in the range of 1 to 10,000 Q-m) are considerably more important than the variations of £r (typically in the range of 1-100). The permeability and permittivity are not important parameters in the context of direct and low frequency current conduction. The dielectric strength of the soil is of importance in phenomena, involving the passage of very high currents in the soil such as lightning surges. Dielectric strength however, remains within fairly narrow limits. This characteristic of earth is analysed in Chapters 5 and 7. We have analyzed the conduction properties of various earth materials. Typically, earth which is investigated by electrical methods is comprised of a combination of a number of materials in varying shapes and sizes and of differing resistivities. Because the conduction of current in the earth is largely electrolytic, the differences in the resistivity values are caused by variations in the amount of, and the nature of the solutes in the water contained in the material. In spite of this usual heterogeneity, there are cases where large, relatively uniform volumes of earth surround a power grounding electrode. A uniform earth assumption is often a reasonable approximation and is a good starting point in helping to understand the fundamentals of grounding. The earth can be broadly modelled as having a voluminous core (3500 km in radius; 2200 miles) consisting of a low resistivity hot magma and surrounded by a 2850 km (1800 miles) thick layer of hot but solid low resistivity material. Another thin layer (50 km below continents; 30 miles; zero to a few km, below oceans) of relatively higher resistivity rock surrounds the thick and hot solid crust. This thin layer consists of strata of different rocks overlain in places by a wide variety of materials. Since large power system ground electrode dimensions are considerably less than the upper crust thickness of earth (0.5 km compared to 50 km), and because the response of ground electrodes is influenced minimally by the nature of soil at distances of more than 10 times their maximum dimension, one may conclude that approximately 1/10 of the uppermost and thinnest earth crust plays a role during current conduction through the electrode. Unfortunately, the widest variations in earth resistivity exist in this surface layer. Figure 4.3 shows an average resistivity map of this upper layer in the United-States. COARSE-GRAINED MATERIAL FINE-GRAINED MATERIAL TEMPERATURE 6 (° C) Figure 4.2 Soil Resistivity Versus Temperature for Typical Soil Samples (Redrawn from [7]) Geologists have measured the apparent resistivity of earth for very large spacings between test electrodes [7,8] (deep penetration of the test current into the soil) and have developed a three-layer earth model, in which the middle layer exhibits a significantly higher resistivity than the adjacent layers, and which gives results in reasonable agreement with measurements. This earth model is reasonably consistent w|th the geological model described previously. The differences are a direct consequence of the distinction which must be made between geological and geoelectric structures. A single geological structure often presents different resistivity values (see Table 4.1) and thus, incorporates several geoelectric structures. The reverse can also occur. Two volumes of different materials may be equal in resistivity and therefore will appear as a uniform geoelectric structure to the geologist. In many cases however, the geological structure will match its geoelectric counterpart. LOW RESISTIVITY < 100 fl-m Figure 4.3 MODERATE RESISTIVITY HIGH RESISTIVITY 100 to 500 fi-m > 500 fi-m Resistivity Map of the United States (Adapted from [7]) While both the geological and geoelectric models are of importance to the geologist, only the geoelectric model is of significance to the power system engineer. Based on the three-layer geoelectric model given previously, the earth can be considered to consist of: 1- A surface layer consisting of several varieties of soil and rock in which a great number of pores or cracks of various sizes are saturated with water containing dissolved salts. The conductivity of this layer is largely electrolytic in nature and its average resistivity is moderate although the actual value varies within wide limits (10 - 10,000 Q-m). The thickness of this layer is in the order of 2 to 5 km (1 to 3 miles). 2- A middle layer, 10 to 40 km thick (12 to 25 miles), characterized by a very high resistivity value (10^ to 10^ Q-m) which results from the considerable pressure impressed on the rocks and the absence of pores and cracks which can retain water. 3- A lowermost layer, consisting of very hot rocks or magma which is highly conductive despite the absence of moisture (10-100 Q-m). This model illustrated in Figure 4.4, is inadequate for modelling power system ground electrodes however, since the significant soil structure surrounding the electrode is represented by a single layer of average resistivity. The thickness of this layer (a few km), although very thin with respect to earth diameter, is nonetheless considerable when compared to the electrode dimensions. To create a suitable earth model for grounding designs, it is necessary to construct one by further subdividing the uppermost layer into additional layers, as required to accomodate the local resistivity values which have the most influence on the response of the grounding system. 4-7 4.2.4 ELECTRICAL MODEL OF EARTH This modelling procedure, which consists of introducing as many horizontal layers of uniform resistivity as necessary, is very attractive since it allows vertical changes in resistivity to be considered as additional layers. In addition, the theory of earth resistivity measurements in multi-layer soils is well understood and solution techniques are available using modern computers. The inverse problem (modelling the earth from measurements) requires considerably more effort and computer time however, and is seldom justified, even for complex geophysical problems. Usually, the computer resources are used in association with empirical methods in order to arrive at a satisfactory approximation. The multi-layer earth model is incapable of recognizing lateral variations in earth resistivity such as vertical faults, or localized volumes of materials with resistivity values different from the surrounding soil. In such cases, suitable equivalent earth models must be derived. Quite often, in practice, various earth models are analyzed separately to approximate a real earth structure. The results are then combined empirically to interpret the measurements. This method generally leads to satisfactory results. For example, a three-layer earth model can be analyzed using the combined results of two different but suitably selected two-layer earth models. This subject will be further discussed in Section 4.5.6. For most power system grounding problems, it has been found in practice [3-5] that a two-layer stratification is a good approximation to the real earth structure, even when measurements indicate that a more complicated structure exists. For the cases where either uniform earth or two-layer models are considered to be unacceptable simplifications, it is necessary to use adjustment algorithms and engineering judgment to predict the performance of a grounding installation. DENSE ROCKS (i n4 - i n 8 n-. Figure 4.4 Geoelectric Model of Earth 4.3 MEASUREMENT METHODS AND TECHNIQUES Because of the wide variations in the structure and properties of earth materials, there are numerous methods and techniques for determining the structure of earth. These methods can be classified into two broad classes: 1- Direct measurements 2- Indirect measurements Direct measurements are based on the extraction, from various depths, of samples of earth materials which are then examined in specialized laboratories. This method is accurate and straightforward provided that the material is not disturbed by the extraction process. However, it requires that expensive borings be made at a large number of representative locations to cover the site being investigated. This kind of investigation is usually performed after indirect measurements have confirmed that the site offers promising results. Such direct geological surveys are used extensively for pre-construction studies. Indirect measurements determine the earth structure from the measurement of the earth response R to certain impressed excitation E. The measured earth response R is a physical quantity which is a function of earth resistivities and a number of other earth parameters. From the known relation between R and earth parameters, the probable earth structure is determined. There are a great variety of indirect methods. These methods can be grouped into the following categories: abcdefgh- Seismic methods Gravitational methods Radio-wave methods Telluric and magneto-telluric methods Induced-polarization methods Induction methods Electrical well logging methods Galvanic resistivity methods. Methods c to h are based on the earth response to electrical excitation sources. A description of all these methods, is beyond the scope of this report. However useful information on them can be found in [7] and [9-15]. The only methods which will be described here are the last two; electrical well logging and galvanic resistivity methods. These tests are used widely throughout North America and typical results are readily available in the literature or from various organizations utilizing the methods. When apparent resistivity measurements (using galvanic resistivity methods) at a transmission or substation site are not available, electric well logs may be available. In the United States and Canada, many hundred of thousands of oil and water wells have been electrically logged. Where this information is not available, civil engineering reports will usually contain borehole log data which can be very useful. This information should be obtained even when resistivity measurements have been performed. By correlating results, the interpretation task will become easier and more accurate. 4.3.1 ELECTRICAL WELL LOGGING Several alternatives exist in the techniques employed and instruments used to measure resistivity. The principles are essentially the same however, regardless of the technique used. Figure 4.5 shows a simplified illustration of the measurement method. A number of small electrodes are lowered into a drilled cylindrical hole. The function of these electrodes is either to inject current into the soil or to measure the potential caused by the flow of electric current through other electrodes. This method is therefore a galvanic resistivity method. It is traditionally considered as a nonconventional method because, the measurements involve a very small volume of soil within the vicinity of the electrodes. Thus, only local changes in resistivity values are recorded. By exploring the well over its length, the vertical earth structure in the vicinity of the well is determined. The differences in the various techniques employed during electrical well logging lie in the arrangement and location of the electrodes. Some commonly used arrangements are depicted in Figure 4.6. 4-9 Power Supply Recorder Lowering Equipment Movable Electrode x (or Array of Electrodes] Figure 4.5 Operating Principle a- SINGLE ELECTRODE LOGS Figure 4.6 Return Current Electrode b- SPACING LOGS c- FOCUSED LOGS d- MICRO LOGS Commonly Used Arrangements A typical resistivity curve obtained from electric logs is shown in Figure 4.7 along with the corresponding geological structure. Accurate interpretation of the apparent resistivity curve requires sufficient expertise to account for the electrode arrangement used and distortion effects caused by the mud in the well. However, an inspection of the curve, based on the knowledge that the apparent resistivity measured is a distorted image of the local volume of earth material, can yield useful information. The interpretation of electric logs has problems similar to those encountered with conventional galvanic methods. However, it is unlikely that this method will ever be used by electrical engineers as a primary measurement technique. Interpreted results will normally be available to the engineer through geological survey reports. Thus, interpretation techniques for this method are not presented. The reader interested in more information should consult [14] 4-10 RESISTIVITY (fi-m) ^tOO 800 1200 DRILL HOLE SECTION Figure 4.7 Typical Log Resistivity Curve 4.3.2 GALVANIC RESISTIVITY METHODS Most commonly used methods for measuring resistivity are based on galvanic contacts between the earth and an array of electrodes. This array typically consists of four electrodes, a generalized arrangement of which is depicted in Figure 4.8. Current is injected into the soil through one pair of electrodes (current electrodes Cl, C2) and the potential established at the surface of the earth by the test current is measured at the second pair of electrodes (potential electrodes PI, P2). As Figure 4.8 suggests, a great variety of arrangements can be used to carry out the measurements. As a general rule however, practical arrangements consist of electrodes driven into soil along a straight line and symmetrically positioned about a central point. Three categories of probe arrangements are generaly used. 1- Potential difference arrangements (measurement of the difference in potential between widely spaced electrodes, PI and P2). 2- Potential gradient arrangements (PI and P2 are closely spaced to measure the local earth gradient). 3- Potential curvature arrangements (both pairs of electrodes closely spaced in order to measure the curvature of the earth surface potential function. All these arrangements can be used to detect vertical or horizontal resistivity changes. 4-11 a - TOP VIEW C2 Figure 4.8 Galvanic Resistivity Methods When vertical resistivity variations are being investigated, the spacing between the current electrodes is increased in steps, thus forcing the test current to penetrate progressively deeper into the lower layers of the earth. Current flow through the earth at greater depths decreases the effect of the surface materials on the potential between PI and P2. Thus, it may be concluded that large spacings between current electrodes are required to determine deep layer resistivities while close spacings are used to determine surface layer resistivities. If horizontal changes in earth resistivity are being explored, the measurement technique consists of moving the entire array with constant electrode spacings along a traverse line. Ideally, this traverse line should be at right angles to the suspected discontinuity which interfaces the two areas of unequal resistivity. The first type of measurement, to locate vertical strata, or layers, is usually designated as "horizontal profiling", while the second type is called "vertical sounding". 4.3.3 THE WENNER METHOD Most resistivity measurement techniques are variations of the four equally-spaced electrode arrangement originally described by Frank Wenner [16]. Variations have been introduced to eliminate or decrease certain difficulties encountered with the Wenner arrangement, particularily at large electrode spacings. Difficulty is caused by very low potential differences which are developed between the potential probes PI and P2 at large electrode separations. Since resistivity surveys for power engineering applications require only moderate spacings, the Wenner method is extensively used. Its simplicity and accuracy are key factors in its popularity. It is essential, however, that appropriate instrumentation be used to mitigate the effect of stray currents in earth. This arrangement, which is of the "potential difference" type, will be reviewed in detail. 4-12 It is more than sixty years since Wenner first described the four probe equal spacing technique for earth resistivity measurements, usually called the Wenner method. Many modifications have been proposed, but the underlying principles are the same. In modern geoelectric surveying, the Wenner method is seldom employed because at large inter-electrode spacings, high sensitivity measuring equipment is required. Moreover, the apparatus used must be capable of discriminating between the measured signal and stray currents in the soil. This problem can be controlled at small inter-electrode spacings by increasing the magnitude of the test current until the noise level is insignificant. In power system engineering applications, the Wenner method is used almost exclusively for resistivity measurements. It is our opinion that this method, when used with suitable test equipment, will provide sufficient data for an accurate earth structure model for the analysis of transmission line structure grounding systems. Because of the relative simplicity of results interpretation and widespread availability of test equipment supporting the use of this method, it is recommended as an effective and suitable standard test procedure. Therefore, the remainder of this chapter will deal almost exclusively with the Wenner electrode arrangement for measuring earth resistivity. The Wenner four-electrode arrangement is shown in Figure 4.9. Four electrodes are driven into the earth along a straight line. The electrodes are uniformly spaced and the burial depth of the electrodes is usually less than 10 per cent of the spacing between two adjacent electrodes. Thus, each electrode will appear as a point with respect to the distances involved in the measurement. A current I is injected into earth via two outer electrodes and the earth surface potential difference V is measured by the two inner electrodes. The outer electrodes are called the current electrodes and the inner pair, potential electrodes. It can be shown that the role of the electrode pairs can be interchanged without changing the electrical response of the array, the ratio V/I. This ratio, with dimension of resistance, is proportional to a variable described as the apparent resistivity pa at spacing a. The proportionality factor between V/I and Pa is called the array geometric factor a . Thus the following equation can be written: (4-3) -if P2 C2 C‘ PI Tijffr a Figure 4.9 a a Wenner Arrangenent 4-13 A similar relation can be written for other arrangements of electrodes. Generally, the geometric factor is different from one arrangement to the other. Interchanging the roles of the electrode pairs does not change the value of the geometric factor. The test current is generally provided from a dc source such as a battery. However, due to polarization at the metal electrodes, a polarity reversal mechanism is used to periodically reverse the direction of current flow between the current electrodes resulting in a chopped dc vaveform. The instrumentation and field measurement techniques are discussed in Chapters 9 and 10. It is also possible to use alternating currents at low frequencies where induction effects and attenuation can be neglected. In such cases, the measured apparent resistivity value is identical to the direct current measured value. Care should be taken to insure that the mutual impedance between test leads is negligible compared to the test electrode ground resistance. This problem will seldom exist at small or moderate inter-electrode spacings. 4.4 DERIVATION OF THE EQUATIONS 4.4.1 POINT SOURCE ELECTRODE IN UNIFORM SOIL Initially, the case of a completely homogeneous isotropic earth and a single point source of current buried at the surface of earth will be considered. The point current source injects a current I in the soil. This current returns to the power source through a return electrode sufficiently remote from the point current source so that its effects can be neglected. The return electrode is assumed to be at "remote" ground. The problem to be analyzed is shown in Figure 4.10. The current source must obey Ohm's law, i.e.: E = pj (4-4) where E is the potential gradient (V/m) J is the current density (A/m2) P is the earth resistivity (fl-m) Since it is assumed that there are no other current sources in the volume of earth, the divergence condition requires that: (4-5) AJ = 0 Combining Equations 4-4 and 4-5 leads to Laplace's equation: AJ =—AE = — v2U P P (4-6) where U is the scalar potential function defined as: 3U=F jU _ 3x x ’ 3y 3U y ’ 3Z z or in a more compact form: U = -grad E (4-7) 4-14 In polar coordinates, the Laplace equation is given by: 9 (/ r 2 -------- )\ +. 9r 9r 9 , 92U . 0 9U x — (sin6 ___) + 90 90 r2sinf r2sin2f 9B2 = 0 (^-8) Because of the complete symmetry of the current flow with respect to the angular coordinates 6 and /?, the derivatives with respect to these variables are zero and (4-8) reduces to: JL (r2 2H) 9r; 9r v = (b-S) 0 SOURCE REMOTE GROUND Figure 4.10 Point Source Electrode in Uniform Soil The solution of this differential equation is straightforward: (4-10) U (r) = - — + B r A and B are constants which are determined from boundary conditions, i.e.: U = / 0 , when r ->- <» Jds = I where s is the surface of a hemisphere concentric to the current source, thus, B = 0 and A is obtained from: /E P dS ~ I r9U Js r2^ dS ~JQ A Pr:2 pr dr = 2 itA p which leads to: A = Pi 2tt consequently: PI U = (4-11) 27rr 4-15 It is possible to obtain Equation 4-11 directly using a hemispherical electrode and simple physical considerations rather than Laplace's equation. Consider a hemispherical electrode having a diameter c and buried at the surface of a uniform earth (Figure 4.11) Figure 4.11 Hemispherical Electrode Because of the spherical symmetry, the equipotential lines are shells concentric to the hemisphere. The voltage drop between two equipotential surfaces located at a distance r and r + dr is: (4-12) IdR = (U + dU) - U = dU where dR is the resistance equipotential surfaces, i.e.: length dR - p of the volume of earth enclosed the two dr - p surface between (4-13) Zirr2 Since dR can be as small as desired, the two equipotential surfaces can both be made equal to ZtitZ. From (4-12) and (4-13) the following equation is derived: pI 1 dU ----------- dr 2ir r2 (4-14) Equation 4-14, integrated between r and «, leads to Equation 4-11. Although convenient for simplified cases, the last derivation can not be generalized to solve more complex problems. In such cases, the first method is required. 4-16 4.4.2 MULTIPLE SOURCE ELECTRODES Since the potential function is a scalar quantity the superposition principle can be applied. The algebraic sum U^ + U2 + ~ + Un of the potentials, established by independent point sources, at a point M is equal to the total potential U which exists when these sources are acting simultaneously. Therefore, the total potential at a point M, due to n point sources at the surface of a uniform earth is: (4-15) M where is the current injected by point source i and rj is the distance between the source and the point M. When Equation 4-15 is applied to the Wenner array, the following potentials are determined at the potential electrodes, (assuming that current at Cl is positive and negative at C2): P I U = ----4iTa p (4-16) PI U p2 4Tia Measured between the inner electrodes, the ratio of the potential difference V, to the test current, I is: V/i = ---- -Sl-L. I p = (4-17) Tra 2 Zira------ (4-18) I Thus, lira is the geometric factor a introduced earlier in this section. This factor is a linear function of the electrode spacing a. Since earth is assumed uniform, a is a linear function of a and the ratio V/I must decrease with increased spacing a. If I is held constant as a is increased, the measured potential V will decrease to a level where stray currents and/or equipment sensitivity will introduce large measurement errors. A straightforward solution would be to increase the value of the test current I. In practice, however, power sources are voltage or power limited and the only feasible way to increase the current is to decrease the ground resistance of the current electrodes. In order to increase the magnitude of the measured voltage V at large probe spacings, one may use the so-called Schlumberger-Palmer array shown in Figure 4.12. This is a symmetrical four probe arrangement with the potential probe moved closer to the current probes than with the Wenner technique. 4-17 Figure 4.12 Sclilumberger-Palmer Array Simple algebraic calculations will show that in this case: ]I The geometric factor a (**"19) of this array is thus r(a2/b - b/4). 7 An increase in the spacing a will not result in a proportional increase of a since b can be increased also. Based on a constant test current I, the measured potential V will decrease more slowly than in the case of the Wenner array. 4.4.3 NONUNIFORM SOILS The concept of a completely uniform earth is useful in defining the apparent resistivity function and its relation to the geometric factors for various electrode arrangements. This concept has also served to establish the fundamental equations which govern the performance of point source electrodes. The analysis of this simple problem can be extended to more complicated earth models, and will serve as a reference for the interpretation of real earth measurements. When resistivity measurements are made at a given site, the measured ratio V/I can be used to define a fictitious apparent resistivity pa as: p V = a(a) a I (4-20) where a is the geometric factor introduced earlier. This geometric factor is a function of the spacing a, when the Wenner arrangement is used. We have shown that a was proportional to a when the earth is uniform. By convention we will assume that a(a) is independent of the earth structure (a(a) = 27ra for Wenner method). The apparent resistivity, so defined, has no physical meaning except when soil is uniform and Pa equals the true earth resistivity. In practice however, the apparent resistivity represents the weighted average resistivity of the earth at the site, up to a depth in the order of the electrode spacing a. 4-18 The apparent resistance R, the ratio of the measured potential difference V to the test current I, is given by Equation 4-17 when earth is uniform. When earth is nonuniform , R is a complicated function of the spacing a and other parameters such as the thickness and resistivity of earth layers: R = ip(a,Ax ,X2................... ,A................... An) (4-21) where Aj is one of the n parameters which define the earth structure. If Paj is the measured apparent resistivity then: Paj " 2Traj^(aj >Ai,A2............... Aj ................An) =0 (4-22) Theoretically, when n measurements are made at various spacings nj, it is possible, to solve the n simultaneous Equations 4-22, to determine the n unknown parameters Xj of earth. This task is not as straightforward as it seems because of several problems which will be discussed in Section 4.5. Before proceeding with this analysis, it is necessary to establish expressions for certain idealized earth structure models which are suitable for the analysis of power system ground electrodes, including transmission line structure grounds. 4.4.4 POINT SOURCE ELECTRODE IN A TWO-LAYER EARTH There are several mathematical approaches which have been used to calculate potentials in a layered soil structure [3,5,17-19]. The method which is adopted here is to search for a particular solution to Laplace's equation which satisfies the boundary conditions of the problem. One of the simplest and most important earth structure models is the two-layer earth model. It can be shown that the potential in earth can be always expressed as the sum of a normal potential (uniform soil), and a disturbing potential, which accounts for the deep layers of soil. Therefore, the two-layer model can be used as the simplest equivalent earth structure for interpreting practical resistivity measurements. We shall analyze the problem of a surface layer of resistivity P} and thickness h, overlaying a second layer of resistivity P2, which extends to an infinite depth (see Figure 4.13). ■ 1 y y' y y s s' s' X ; Tjjjp h ^ p1 wv\ r z M(x,y,z) ipfj 1 ^2 1 Figure 4.13 Two-Layer Earth Model 4-19 The potential U at point M must satisfy Laplace's equation. Since there is cylindrical symmetry, Laplace's equation reduces to: 3ZU 1 3U 32U 3r2 r 3r 3z2 ----- +-------- + ------ = (4-23) 0 Assume now that a solution can be found in the form of a product of two independent, single variable functions namely: U = 0 (r) 4>(z) Substituting U in Equation 4-23 and rearranging, leads to 32i|j ------- = 3z2 320 1 (4-24) 0 30 ----- + -—-+a20 = O 3r2 r 3r (4-25) The solution of (4-24) is >p(z) = A(ct) e°Z + B(o) e~OZ The solution of (4-25) contains terms with Bessel's function of zero order of both the first and second kind 0(r) = C(a) J0 (or) + D(ct) Y0 (ar) The function Y0(<rr) can not be retained in this case, since it tends to become infinite when r is small. Consequently, its coefficient D must be set identically to zero. The solution of Laplace's equation is therefore: X ' y u = 2tt >[3(a) Jo(ar) + 0(a) J0(ar) e°z (4-26) L where Pil 3 =---- A C 2tt PjI and 0 = ------ B C 2fT Since any linear combination of terms in (4-26) is acceptable, the general solution of Equation 4-23 may contain an infinite number of terms with different values of a. Since a is arbitrary and thus, can vary continuously, the general solution must be given by: U 3(a) Jo(ar) eOZ da + 4-20 J0 (ar) -crz da (4-27) The values of 0 and constants are determined from the boundary conditions: 9 1- The potentials of the upper and lower layers must be equal at the surface boundary of the two layers: Ux(z) = U2 (z) ; z = h (4-28) 2- At the interface between the two-layers (z = h), the normal component of current flow must be continuous: 1 3Ux 1 3UZ ----------- = _ ----pi 3z p2 3z ; (4-29) (z = h) 3- No current can flow accross the upper layer and air boundary: 3Ui ----3z ; =0 (z = ) (4-30) 0 4- As z goes to infinity, the potential must go to zero: z -s- co Uz(z) = 0 ; (4-31) The upper and bottom layer potentials and U2 are given by Equation 4-27 where /3 and 6 are replaced by j8x, #1 and 182, @2 respectively. When Ui and U2, as given in (4-27), are replaced in the boundary Equations 4-28 to 4-31, a set of 4 equations with four unknowns, namely, @1, 02, 9y and #2 are obtained. The solution of the simultaneous equations is straightforward but lengthy and has been omitted. Since all the integrations are carried out through the same limits, the integrands are identically zero when the integral is equated to zero. The following results are obtained: 1 0i (a) 1 -ke ke (a) = -2ah 1 -2ah - ke -2ah (4-32) + k -2ah 1 -ke 1 62(a) ; 62(a) = 0 where k is designated as the reflection coefficient: p2 - Pi K = ------------P2 + Pi (4-33) The fraction: 1 , - ke-20h can be expanded as: 1 -ke 1 -2crh - y>n« -2nat) (4-34) n=1 4-21 noting that: — = (r2 + z2) ^ = I e az J0(ar)da r1 J0 and then replacing and 0^ by their expansions as given in (4-34), into Equation 4-27 and integrating, the following expression for the potential at a point M on the earth surface (z=0) is obtained: U: = PLifi 27rr (^-35) C1+(2nh/r) Thus the potential consists of two parts: a- The normal potential (PiI)/(2 tit) which exists if the soil was uniform (infinite layer of resistivity P i). b- A disturbing potential caused by the presence at finite depth of a layer having a different resistivity P 2 than the upper layer. 4.4.5 POINT SOURCE IN EARTH WITH INCLINED LAYERS If the boundary separating two regions of earth with different resistivities is not horizontal, but inclined at an angle to the surface, an exact mathematical solution for the potential function is arduous to obtain. The potential function is obtained through a double integration process on two dummy variables, of a complicated function containing hyperbolic sines and Bessel functions of the second type [3,20,21]. When the angle of dip is 90°, a simple solution can be obtained. Figure 4.14 illustrates this case. The case where a vertical interface between two dissimilar earth resistivity areas exists, will be referred to as a vertical fault structure. In practice, this effect can be the result of a number of different mechanisms, including, but not limited to, geological faulting. S I I I f b - TOP-VIEW a - SIDE-VIEW Figure 4.14 Z Vertical Fault 4-22 3 The potential about a current source embedded in earth at a distance d from the boundary separating two media of different resistivities Pi and P2, may be found by an analysis which follows the same lines developed for the case of two horizontal layers. Equations 4-23 to 4-27 are also applicable in this case. By transforming the original semi-infinite medium into an infinite medium symmetrical about the earth surface plane, the potential function U given by (4-23) must satisfy the following boundary conditions: a- Ui must go to zero as z goes to - » b- U2 must go to zero as z goes to 00 c- At the surface between the two media (z=h): U1 (z) = U2(z) d- At the surface between the two media (z=h): 1 SUj Pi 9z 1 pa 9U2 9z These boundary equations are suffieient to determine the unknown constants 6^, 82 and 02* Simple algebraic manipulations will show that the potential at any point of the earth surface is given by one of the following expressions, depending on whether M is in region 1 or 2. (4-36) (4-37) If the current source is embedded in region 2, the above expressions are also applicable provided that Pj and P2 are interchanged. 4.4.6 POINT SOURCE EMBEDDED IN A LOCALIZED DISCONTINUITY Thus far, we have analyzed the case of a nonuniform earth consisting of two layers extending to infinite distances. In practice, there are circumstances where the current source is buried in a small volume of material having a resistivity significantly different from the average earth resistivity of the site considered. In such case, neither the horizontal nor the vertical layered earth models are suitable to model the real situation. Examples of these cases are electrodes buried in a small lake, or electrodes in a small volume of soil artificially treated to lower resistivity. This kind of problem is easily solved if it is assumed that the current source is embedded at the center of a hemispherical volume of soil of resistivity Pj different from the resistivity P2 of the rest of the earth (see Figure 4.13). 4-23 Figure 4.15 Localized Discontinuity This earth structure has spherical symmetry and therefore Equation 4-10 is applicable: U = - ^ + B (4-38) The boundary conditions are: Ui = U2 j_ 3U Pi 3r U2 “ ; r = R _ J__ 9U . p2 3r ’ when r 0 (4-39) r R (4-40) (4-41) 00 X' pi Jids = I The solution of Pil Ux =----2ir (4-42) Equations 4-58 to 4-42 leads to: 1 L r Pz~Pi (4-43) pi R Pzl U2 = (4-44) irr 2 It should be noted that once again, the potential Uj in the upper layer consists of two terms, the normal potential and a disturbing potential. In this case however, the "disturbance" is constant regardless of the location of the point M. 4.4.7 APPLICATION OF THE POTENTIAL FUNCTIONS TO THE WENNER ARRAY The apparent resistivity pa, as measured using the Wenner method, is obtained easily from the surface potential function previously derived for various earth structures. The Wenner arrangement consists of two current sources of opposite polarities separated by a distance 3a, and two potential probes, located between the current electrodes at distance a from 4-24 each other and from a current source. Applying the superposition principle to a group of current sources, the apparent resistivity pa is determined as (see Figure 4.16): + U Pa +1 Cl 1 Cz t i i • P2 ♦pi i a (Pj) - U Cl (P2) (4-45) -I C2 i a 2 Figure 4.16 Wenner Array Two-Layer Earth (Horizontal Layers) From Equations 4-35 and 4-45, the following expression is obtained for pa; Pa = Pi , . co 'Ey? n=1 r [ ---------L (4-46) L [1+((2nh/a)2r 2nl [4+(2nh/a)J. Vertical Fault (Vertical Layers) Equations 4-36, 4-37 and 4-45 are used to derive the expression for Pa. In this case however, several expressions are necessary for each location of the center of the array O whose direction is at an angle « to the line of fault (see Figure 4.17). Line of Fault Figure 4.17 Vertical Fault (Top View) 4-25 Let M be the center of the array and O the point of intersection between the array direction and the line of fault. The distance between M and the line of fault is h. The following geometric relations exist between h, « and the distances dj and 62 between the current sources and the line of the fault: : S COSO) (4-47) s sinoi (4-48) 1 3 = h + — a s 1 nto (4-49) 2 3 = h-----a (4-50) s t no) 2 By convention, we will assume that Cl is the positive current source, located in the medium of resistivity p^. Furthermore, the center of the array is also assumed to be in region 1. This arrangement is symmetrical with respect to the vertical line OX. The results which are obtained using this assumption will also be applicable in region 2 by interchanging and P2 in the equations. Therefore there are three possibilities: 1- P2, C2 in region 2 2- P2 in region 1, C2 in region 2 3- P2, C2 in region 1. In the first case, we have: U(Pi) = UnCPj) + U2i (Pi) (4-51) U(P2) = U12(P2) + U22(P2) (4-52) where Uj; is the potential at a point in region j due to a source in region i. From Equations 4-35 , 4-37 , 4-45 and 4-47 to 4-52, the following expression is determined: Pi PA = --------3 K(1-K) 1 K(1+K) (4-53) + k2 + (1-k) [4(sinw+h/a) 2+cos2w] 2 [4(s inoj-h/a) 2+cos2(jo] 2. Similar expressions can be obtained for the other cases. Localized Discontinuity In Section 4.4.6, the derived potential expression applies only when the current source is at the center of the hemispherical discontinuity. In a typical resistivity measurement array, a current source is unlikely to be located at this center. In this case, the potential calculations are significantly more difficult. It is possible however, to obtain an approximate solution for the apparent resistivity function by initially considering a combination of the following cases: a- current source located far from the discontinuity b- current source located close to the discontinuity center and then, through appropriate interpolations, completing the apparent resistivity function. 4-26 4.5 INTERPRETATION OF THE MEASUREMENTS The information contained in this section will provide a base of knowledge from which the results of apparent resistivity measurements can be correctly interpreted. There are a variety of interpretation methods and techniques used by geologists. Not all of these methods are applicable to power system grounding problems. The methods which we believe to be most useful to power system engineers are presented and discussed in detail. 4.5.1 BASIC CONSIDERATIONS The simplest interpretation problem is the case where the measured apparent resistivities Pa vary minimally around an average value p. This indicates that earth at the measurement site is reasonably uniform and has a resistivity p. Typical apparent resistivity curves peculiar to this situation are shown in Figure 4.18(a). SPACING a- QUASI-UN I FORM SOIL Figure 4.18 SPACING b- ERRATIC MEASUREMENTS SPACING c- THIN UPPER-LAYER Uniform Earth Observed resistivity variations can be attributed to small local discontinuities, which may be neglected, or to inaccuracies in the measurements due a number of factors such as stray currents in earth or inadequate sensitivity of the measuring equipment. Unfortunately, such cases rarely occur in practice. In most cases, apparent resistivity, plotted as a function of the electrode spacing shows large variations with probe spacing. This indicates that the earth is nonuniform. In general, apparent resistivity curves change smoothly and do not exhibit sudden changes. When the latter occurs, it is a clear indication that the array has just crossed a vertical fault or a local discontinuity close to earth surface. The magnitude of the jump is an indication of the difference between the resistivities of the two adjacent earth materials. The presence of buried pipes or other structures close to the surface is a typical cause of sudden changes in apparent earth resistivity (see Figure 4.19). 4-27 SPACING Figure 4.19 a Sudden Changes The method used for interpreting the measurements can be grouped into two simplified categories: • Empirical interpretation • Analytical interpretation Analytical interpretation is, in theory, independent of the person conducting the interpretation. In contrast, the results of an empirical interpretation are significantly influenced by the background and experience of the interpreter. It is preferable to use a combination of both approaches for maximum accuracy and a minimum of uncertainty. For example, when analytical methods indicate that two or more earth models are reasonable, the most realistic choice can be determined from empirical considerations or visual inspection of the curves. In any case, it should be emphasized that experience is of paramount importance in the interpretation process. 4.5.2 EMPIRICAL METHODS Empirical methods are based on experience gained through numerous measurement and interpretation exercises. Thus, such methods can be described as statistical in nature. Essentially, it is observed that the shape of an apparent resistivity curve is closely related to the earth structure and its characteristics at the site. Therefore, certain properties of the measured curve are used to deduce the resistivity and thickness of the earth layers. Although there may be inherent inaccuracies in some of these methods, they are of less consequence to the design engineer than they are to the geologist. Empirical methods may be useful for on site interpretations and serve as a good starting point for more rigorous methods. 4-28 The Gish and Rooney Rule Based on their work, Messrs Gish and Rooney [22J concluded that apparent earth resistivity measurements give the average resistivity of earth material to a depth approximately equal to the electrode spacing. Based on this assumption, a change in the formation can be detected quickly when a significant variation of the apparent resistivity value or a change in the curvature of the data is observed. This rule implies that the test current does not penetrate to a depth greater than the electrode separation. This is incorrect as can be seen from Figure 4.20 which shows the magnitude of the current density as a function of depth, computed for the case of uniform soil. In soil with a low resistivity bottom layer, the current densities in this layer are even greater. 3 $ a Figure 4.20 H • Current Density in Earth Consequently, the empirical rule established by Gish and Rooney should be used with caution as it may lead to false conclusions, and is, at best, a coarse approximation. The Lancaster - Jones Rule Quite often, the apparent resistivity curve indicates the presence of two major electrically distinct layers. These layers are described as electrically distinct to distinguish them from the geological structures which may be comprised of more than two distinct layers. In such cases, the apparent resistivity curve contains a point of inflection somewhere between the small spacing values, where the upper layer resistivity value primarily influences the measured values, and the large spacing values, where the bottom layer resistivity primarily influences the measured values. Messrs Lancaster-Jones [23] have suggested that this point of inflection occurs when the spacing between the electrodes in the Wenner array is approximately equal to 3h/2, where h is the thickness of the upper layer. 4-29 While the theory of two-layer earth shows that there is a relation between the point of inflection, the thickness h and the spacing a, it also shows that this relation is dependent on the reflection factor K (i.e., contrast between the upper and bottom layer resistivities) and varies between two close limits. Only the upper limit is close to 3h/2 and corresponds to the case where the bottom layer is significantly more resistive than the upper one. Based on Tagg's work [3], it appears that a more appropriate relation between the spacing a and h at the point of inflection is: a = (0.7K + 1)h, when K > 0 (h-5b) a = (0.15K + 1.15) h , when K < 0 (4-55) and It should be pointed out however, that the exact location of the point of inflection is not easily determined by visual inspection. This difficulty further reduces the accuracy of the Lancaster-Jones Rule or its improved version. Nevertheless, this rule can be useful in many cases, particularly whe'n the knowledge gained is used to select the most suitable location for the next resistivity traverse. The Asymptote's Rules This method is applicable when the bottom layer is significantly more resistive than the upper layer. This interpretation procedure is illustrated in Figure 4.21 which shows the results of a typical apparent resistivity measurement. Note that the resistivity curve is usually drawn on bilogarithmic graph paper. This method [7] is based on the knowledge that an apparent resistivity curve will rise at a 45° angle if the deep layer is an insulator and if the spacings are large enough. First, a horizontal line (horizontal asymptote), passing through the measured point at short spacings, is drawn. The point of intersection with the resistivity axis gives the upper resistivity. Secondly, a 45° line (asymptote) passing through the last measured point(s) is drawn. The ratio of spacing to apparent’resistivity for any point along the 45° line, C, is constant and is used to determine the thickness of the upper layer. The value of C/(21n2) represents the columnar conductance, in ohms, of the upper layer. The thickness of this upper layer, when the Wenner arrangement is used, is: C a ------------ = --------------px 21n (2) 2paln{2) (4-56) The resistivity of the bottom layer is indeterminate, since this method is valid only if the bottom layer resistivity P2 is much greater than the top layer resistivity Pj. In practice, this means that P2 can be anywhere between approximately 50 P^ and 00. 4.5.3 ANALYTICAL METHODS The word "analytical" can be misleading as it is often interpreted as meaning "accurate" or "rigorous". Earth resistivity measurements are rarely accurate to within 1%, even when sophisticated equipment is used. Usually, careful measurements with conventional equipment are accurate to within about 5%. Careless measurements, inexperience or poor equipment can lead to measured results significantly different from the real values. Even measurements taken by an experienced crew under the best of conditions, will never give a perfect match with analytical results computed from the optimum earth model derived from the measurement data. The reasons for this were briefly explained in the introduction and in Section 4.4.3. We will now elaborate on these effects. 4-30 >00 00 LU C£ CC < CL. CL. < SPACING Figure 4.21 Asymptote's Rule It was shown in Section 4.4.3 that the relation between the measured apparent resistivity Paj at spacing aj is given by Equation 4-22 which is now rewritten: PaJ- = ZtiajiMaj > > *2 ........... -X.----- ------ »An) (4-57) Xi,....,Xn are n parameters which completely describe the real earth structure and its characteristics. In theory, n measurements at different spacings are sufficient to uniquely solve the n unknown parameters of the n simultaneous equations given in (4-57). Unfortunately not only the values of the parameters are not known, but their nature, number and influence on the measurements are also unknown. Since one can not see the various materials contained in the earth, it is not possible to define which variables describe the shape, location and resistivity of the material. One possible solution to this difficulty is to consider the resistivity of earth as a function of space, i.e., P(x,y,z). Unfortunately, even if an analytical solution to this problem was possible, it would not be very useful, since the real function is generally discontinuous at the interface between two volumes of different materials. This returns us to our first difficulty, which is to determine the location and shape of the boundaries where the resistivity discontinuities exist. It is in this determination where the fundamental difficulty associated with earth resistivity measurement lies. It is necessary to propose an earth model and see whether or not the computed apparent resistivities of the proposed model fit those measured. It is obvious that a perfect match will seldom occur. Even if satisfactory agreement is obtained between the measured and calculated results, measurement inaccuracies always make it possible that results from another earth model could be matched equally well with the measured results. 4-31 4.5.4 METHODOLOGY Based on the discussion of the preceding section, it can be concluded that what we have termed "analytical methods" follow a constant methodology, i.e.: Step 1 The measured results are examined and preliminary interpretation is performed, typically based on the empirical methods described previously. Step 2 One or several possible earth models are proposed. Step 3 The measured results are compared with those calculated from the proposed models. Step 4 The most suitable model is retained. If more that one model is suitable, these are considered to be equivalent. Step 5 The selected model is optimized. Often the optimization process is based on engineering judgment. Sometimes one may choose to conduct additional surveys in order to check the validity of some assumptions or to eliminate uncertainties. However, in power system design this step is generally omitted. The differences in techniques employed to interpret the results are essentially due to: 1- Complexity of the earth model selected. 2- Availability of apparent resistivity Master reference curves based on the proposed model. 3- Techniques used to make the comparisons between measured and calculated values. Two different techniques are used to match calculated and measured results: a- Complete-curve matching b- Partial-curve matching. In some cases, prepared apparent resistivity curves are used to compare measured and calculated results. When these reference curves are not available, computer programs are normally used. There are many difficulties which, in practice, limit the complexity of the earth model to a small number of layers. In this report, only the two-layer earth model is considered in detail. Three or more layers are studied briefly using the technique of partial-curve matching for two-layer earth models. First, the use of precalculated curves is described. Computer methods based on the method of steepest-descent are then described. A computer program (RESIST) based on the method of steepest-descent is described in Chapters 3, 10 and Appendix B. 4-32 4.5.5 LOGARITHMIC CURVE MATCHING The apparent resistivity functions which were derived in Section 4.4.7 (Equations 4-46 and 4-53) may be written in terms of the dimensionless ratios K, Pa/ and h/a: Horizontal Two-Layer Earth CO = 1 + 4 ^ ' Kn^[l + 4n1 2 (h/a)2] 2 - [4 + 4n2 (h/a)2] 2] (4-58) P-/Pi n=1 Vertical Fault j/Pj - (l_K) 1 + K2 + K(1-K) [ 4(sino) + h/a)2 + - K(1 + K) [4(sinoj - h/a)2 + Cos2oj] Cos2oj 2 J ] 2 (4-59) If one of the above dimensionless apparent resistivity functions Pa = pa/is plotted in logarithmic coordinates, i.e., ln(Pa/ P^) = F(ln(a/h)), the coordinates of a point on the resistivity curve will be: Y = lap a - 1npi x = 1na - Inh on the y axis (4-60) on the x axis (4-61) If we now assume that a number of apparent resistivity reference curves, designated as Master curves, are plotted for various reflection coefficients K, then: y° = Inp (4-62) 3 x° = Ina (4-63) assuming that Pi = 1 fi-m and h = 1 m. Where Pi and/or h are not equal to unity, resistivity curves derived for two-layer earth structures will be shifted by -ln(Pi) vertically and -ln(h) horizontally, with respect to the corresponding Master curves as shown in Figure 4.22. The shapes of the curves are thus preserved. This property of the apparent resistivity curve is the basis for the logarithmic curve matching method. Thus, a field curve can be compared directly with a set of theoretical Master curves, through a series of appropriate translations of the field curve plotted on transparent paper. If a satisfactory match is found between the field and a theoretical curve, then the real earth reflection factor K, is equal to that of the computed curve. Pi and h are then determined as follows: 1- Pa/ Pi = 1 on the Master chart, corresponds to 12 fi-m on the field chart. Therefore: 12/Pi = 1 and Pi = 12 fi-m 2- a/h =1 on the Master chart, corresponds to 30m on the field chart. Thus: 30/h = 1 and h = 30 m. 4-33 ^__ L * t_ MEASURED VALUES HORIZONTAL t* VERTICAL Figure 4,22 Logarithmic Curve Matching This method requires that a set of precalculated Master reference curves be available to the interpreter. Such curves can be easily determined based on Equation 4-58 for horizontal layers and Equation 4-59 for a vertical fault. Figure 4.23 is a set of Master curves for the case of a two-layer earth. When a field curve falls between two curves, the correct values can be interpolated. If more precision is required, additional curves for closer K values should be constructed. Figure 4.24 is a set of Master curves applicable to vertical faults when the direction of the traverse is at 0° angle with the line of fault. It should be noted that additional charts (different angles w) are required for a complete set of master charts. The construction of such charts is straigthforward with a programmable calculator. Reference charts for 30, 60 and 90° angles are provided in Volume 2 of this report. At this point, a word of caution is necessary. The apparent resistivity curve measured when a traverse does not cross the line of a fault (for example, when the direction of the measurement array is parallel to a line of the fault) is very similar to a typical horizontal layer curve. In practice, it may be very difficult to distinguish between the two situations. This difficultiy can be readily overcome if two traverses, at right angles, are made available. In the case of horizontal layers, no significant differences will be found between the two traverses. If the earth exhibits vertical discontinuities, the curves of each traverse will be different. The differences may be significant if the vertical fault is close to the traverses and/or the resistivity difference between the regions is marked. If one of the traverses is more or less orthogonal to the fault and crosses the discontinuity, then one curve will have a shape similar to the curves shown in the figures of Volume 2 (w = 60° or 90°). TWO-layer Figure 4.23 earth Horizontal Two-Layer Earth Master Chart 4.5.6 PARTIAL CURVE MATCHING The logarithmic curve matching method is a complete curve matching technique. That is, the best fitting theoretical curve is used as a reference to calculate the unknown parameters of the real earth. The occurence of more than two-layers in earth is quite common. In many cases, however, it is necessary to average the apparent resistivity values in the zone which indicates the presence of a middle layer and use the new adjusted curve as the reference field curve (see Figure 4.25). 4-35 9£-*? VERTICAL FAULT 10 Figure 4.24 Vertical Fault Master Chart -5 SPACING SPACING Figure 4.25 Multi-Layer Earth Unfortunately, if the difference in resistivities between adjacent layers is significant, this approximation may be unacceptable. Even though multi-layer earth problems are not analyzed here, it is necessary to evaluate the nature of multiple layer structures in order to choose the best approximate two-layer model. The partial curve matching technique is recommended as appropriate for this evaluation. This technique, illustrated in Figure 4.26, is essentially a complete curve matching process, conducted in two (or more) consecutive steps on given zones of the apparent resistivity curves. Connecting Zone SPACING Figure 4.26 Partial Curve Matching 4-37 Once the earth model for each portion is determined, the models are overlaid to arrive at a suitable multiple layer earth model. It is not practical to develop rules to define the resistivities or the thicknesses of the layers. Engineering judgment and experience must be the major guiding factors. A typical example of this technique is described in Chapter 10. Alternatively, partial curve matching can be performed using a two-layer resistivity interpretation computer program, simply by submitting the data for zone 1 to the computer followed by the data for zone 2. The rest of the interpretation task is identical whether Master curves or computer programs are used. In grounding problems where a maximum of two-layers are considered, one more step is necessary. The three (or multi) layer earth model derived using the partial fitting technique, must be converted to an equivalent two-layer model by combining the effects of adjacent layers. The major difficulty is in the selection of those layers which can be combined without seriously affecting the accuracy of the grounding analysis. If we restrict ourselves to the case of three layers, then the main problem is whether the mid-layer should be combined with the upper or bottom layers. Theoretically, the answer will depend on a number of factors such as the ratio of the grounding system dimension to the upper and mid-layer thicknesses, the contrast between the resistivities and the presence of ground rods in the mid or bottom layers. Generally however, it is preferable to combine the mid and bottom layers because, as explained in Chapter 10, the surface potentials above the grounding system are particularly sensitive to the characteristics of the upper layer, which usually encloses the ground conductors. The influence of the deep layers should not be significantly different whether the deep layers are represented discretely or as a lumped equivalent layer. This observation is also applicable to the calculation of resistance value of the grounding system [24-26]. Similar observations and reasoning can also be applied to multi-layer earth. 4.5.7 OTHER METHODS There are many other methods which have been proposed and used to interpret apparent resistivity curves. These methods use graphical or computation techniques based on the known properties of idealized earth models [27-29]. Tagg has proposed two techniques which could be adapted to either graphical or computerized approaches [3]. Most computerized methods are essentially of the curve fitting type using a known reference function. The selection of a suitable curve fitting technique is based on considerations such as speed of computations, stability of the algorithm, convergence and ease of programming. The following method was used as the basis for the computer program RESIST (see Appendix B) developed by SES. This method was selected for the following reasons: • A more elaborate version of the method has been operational for several years and has been extensively tested in practical cases. It has proved to be a very stable, reliable method which requires a minimum amount of input data. Initial values for the presumed earth model are not required from the user. Therefore, it can be used by engineers with limited experience in resistivity interpretation. • The method is easy to implement • The program is particularly suited to transmission line structure grounding design where very large spacings between the electrodes of an array are not necessary. This program is based on the method of steepest-descent described in the following section. 4.5.8 THE STEEPEST-DESCENT METHOD General The method of steepest-descent [30] is most readily vizualized by an analysis of the two-variable function ^(x,y), illustrated in Figure 4.27. The gradient of this function is calculated at an initial point M0 defined by x0, y0. The values of x and y are then selected so that the function decreases along the direction defined by the gradient vector. The process is repeated until the function along the initial direction starts to increase. The process will stop when all possible directions of the gradient indicate that the present (x,y) coordinates corresponds to a minimum of the function (zero gradient). MINIMUM Figure 4.27 The Method of Steepest-Descent This process will normally converge to a minimum of the function. However, there is no guarantee that the minimum obtained will be the only one nor that it is the minimum of the minima. Experience shows that when a secondary minimum is obtained, the initial starting point was most likely within the zone of influence of this minimum. In this case another pair of initial x and y values should be selected and the process started again. Analytical Description Let p°(aj), j = 1, n be the series of apparent resistivity values as measured at a given by the Wenner method for n different inter-electrode spacings aj. site Let p(ap, j = 1, n be the calculated apparent resistivity value, based on a two-layer earth model and the same spacings aj used during the measurements. The interpretation task consists of finding the most suitable earth model for which the difference between the set of measured and calculated values, according to certain criteria, is a minimum. In theory anycriterion can be used (e.g., sum of the absolute value of the differences). In practice, the classical least-square criterion is preferred. 4-39 Let ^(Pi, K, h) be the square error function defined as: 'P°(ai)~p(al)' (it-64) ijj(pi, K, h) = j=1 L PUU., The best fit is obtained when ^ is minimum. The values of Pi, K, h which lead to this minimum are determined by the steepest-descent algorithm. The gradient vector is defined as: / 8^ 8^ 3^ \ \ 3pi ’ 3h ’ 9K / Each component of the potential vector is determined from Equation 4-64. Thus: 8i|) - 3pi dip 2 3p E(pl'p) 1 V p02/ 3pi - - 2 t(p0-p) 3p 3^ 3p K > O 1 \ P02/ 3K ! 3K II (4-65(a )) (4-65(b)) (4-65(c)) 1 \ p02/ 3h 3h Assume now that represented as: AP^,, AK, Sip Api = (4-66 (a)) -T ------- Spi 3ip AK = (4-66(b)) -T -------- 3K Sip Ah = -x---3h (4-66(c)) where r is a positive value (expressed in p.u. of V), suitably selected to generate a smooth search for the minimum. The above changes cause a small variation in the error function A<J> =-g- Api +4r"AK +lh~Ah or (4-67) 4-40 The sought for minimum is obtained when = 0 or practically when: |Ai|»| < e (4-68) where e is the desired accuracy. The main steps in the steepest-descent algorithm are therefore: 1- Estimate initial values of K and h (i.e., Pj , K° , h° ) 2- Calculate a suitable value of r 3- Determine APj, AK and Ah 4- Estimate a new starting point: + Api K<.) h<'> + AK hC-D + Ah 5- Calculate [A'/'] and compare it with e: a- if | A^| < e , the fit is completed. b- if |A^| > e , continue the process at step 2 (or step 3 if maintained constant). t is In order to calculate A\4 from Equation 4-67, the partial derivatives of \p must be known. These are determined from (4-65) where the partial derivatives of the theoretical two-layer earth apparent resistivity function are obtained from Equation 4-46. These calculations lead to: 3p_ (1 - n(1 - K2)/2K) (A" (4-69) 9pi CO 3p ^ nKn_1 - B-i ^ (4-70) 3K 9p 3h a2 (4-71) n=1 where A = 1 + (2nh/a)2 (4-72) B = A + 3 4-41 REFERENCES 1 - B. L. Goodlet, "Lightning" IEE Journal, Vol. 81, No. 487, July 1937, pp. 1-56. (see discussion by F. J. Miranda). 2 - A. Cimador, R. Fieux, B. Hutzier, "Influence des Resistances des Prises de Terre sur le Foudroiement des Pylones d'une ligne 222 kV", E.D.F. Bulletin de la Direction des Etudes et Recherches, Serie B, No. 4, 1974, pp. 29-42. 3 - G. F. Tagg, "Earth Resistances", George Newnes Ltd., London 1964 (book). 4 - E. D. Sunde, "Earth Conduction Effects in Transmission Systems", Dover Publications, New York, 1968 (book). 5 - F. Dawalibi, D. Mukhedkar, "Influence of Ground Rods on Grounding Grids", IEEE Transactions on PAS, Vol. PAS-98, No. 6, November/December 1979, pp. 2089-2098. 6 - P. Hoekstra, D. McNeill, "Electromagnetic Probing of Permafrost", Geophysics 1973, pp. 517-526. 7 - G. V. Keller, F. C. Frischknecht, "Electrical Methods in Geophysical Prospecting", Pergamon Press, New York (1977). 8 - G. V. Keller, "Statistical Study of Electrical Fields From Earth Return Tests in the Western States and Comparison with Natural Electrical Fields", IEEE Transmission on PAS, Vol. 87, April 1968, pp. 1050-1057. 9 - G. E. Backus, J. F. Gilbert, "Numerical Applications of a Formalism for Geophysical Inverse Problems", Geophysics J. R, Astr. Soc.13 (1967), pp. 247-276. 10- L. C. Cham, D. L. Moffat, L. Peters Jr., "A Characterization of Subsurface Radar Targets", Proceedings of the IEEE, Vol. 67, No. 7, July 1979, pp. 991-1000. 11- M. Cauterman, J. Martin, P. Degauque, R. Cabillard, "Numerical Modeling for Electromagnetic Remote Sensing of Inhomogeneities in the Ground" Proceedings of the IEEE, Vol. 67, No. 7, July 1979, pp. 1009-1015. 12- V. I. Dimitriev, M. N. Berdichevsky, "The Fundamental Model of magnetotelluric Sounding" Proceedings of the IEEE, Vol. 67, No. 7, July 1979, pp. 1034-1044. 13- H. O. Seigel, "The Induced Polarization Method", Mining and Groundwater Geophysics, 1967, Published by the Geological Survey of Canada, pp. 123-137. 14- H. Guyod, "Interpretation of Electric Ray Logs in Water Wells", The Well Log Analysis, January-March 1966. 15- L. B. Slichter, "An Inverse Boundary Value Problem in Electrodynamics", Physics, Vol. 4, December 1933, pp. 411-418. 4-42 16- F. Wenner, "A Method of Measuring Resistivity", National Bureau of Standards, Scientific Paper 12, No. S-258, 1916, p. 499. 17- S. Stefanesco, C. & M. Schlumberger, "Sur la Distribution Electrique Potentielle Autour d'une Prise de Terre Ponctuelle dans un Terrain a Couches Horizontales Homogenes et Isotropes", Journal de Physique et Radium, Vol. I, Serie VII, No. 4, 1930, pp. 132-140. 18- M. Muskat, "Potential Distribution About an Electrode on the Surface of the Earth", Physics, Vol. 4, No. 4, April 1933, pp. 129-147. 19- H. M. Mooney, E. Orellana, H. Pickett, L. Tornheim, "A Resistivity Computation Method for Layered Earth Models", Geophysics, Vol. XXXI, No. 1, February 1966, pp. 192-203. 20- K. Maeda, "Apparent Resistivity for Dipping Beds, Geophysics, Vol. XX, No. 1, January 1955, pp. 123-139. 21- L. S. Palmer, "Examples of Geoelectric Surveys", IEE Journal, Vol. 106, part A, June 1959, pp. 231-244. 22- O. H. Gish, W. J. Rooney, "Measurement of Resistivity of Large Masses of Undisturbed Earth", Terrestrial magnetism and Atmospheric Electricity, Vol. XXX, p. 161. 23- E. Lancaster-Jones, "The Earth Resistivity method of Electrical Prospecting", The Mining Magazine, June 1930. 24- F. Dawalibi, D. Mukhedkar, "Parametric Analysis of Grounding Grids", IEEE Transactions, Vol. Pas-98, No. 5, September/October 79, pp. 1659-1668. 25- F. Dawalibi, D. Mukhedkar, "Influence of Ground Rods on Grounding Grids", IEEE Transactions, Vol. PAS-98, No. 6, November/December 79, pp. 2089-2098. 26- J. Endreyni, "Evaluation of Resistivity Tests for Design of Station Grounds in Nonuniform Soil", IEEE Transactions on PAS, Vol. PAS-84, No. 12, December 1963, pp. 966-970. 27- L. B. Schlichter, "The Interpretation of the Resistivity Prospecting Method for Horizontal Structures", Physics, Vol. 4, September 1933, pp. 307-322. 28- A. F. Stevenson, "On the Theoretical Determination of Earth Resistance from Surface Potential Measurements", Physics, Vol. 5, April 1934, pp. 114-124. 29- A. A. R. Zohdy, "Automatic Interpretation of a Resistivity Sounding Using Modified Dar Zarouk Functions" Geophysics 38, pp. 644-660. 30- F. Scheid, "Numerical Analysis", Shaum's Outline Series, McGraw-Hill Book Company, New York 1968. CHAPTER 5 POWER FREQUENCY PERFORMANCE OF TRANSMISSION LINE STRUCTURE GROUNDS 5.1 GENERAL Ideally a power system should not conduct any ground currents when only one energized element in the system is grounded. In practice, however, all elements of a power system have a capacitance to ground due to the finite distance between each element and the earth's surface. Therefore, there is always an earth return path for the current through the distributed capacitance of the power system during normal operation. The earth acts as a return conductor during faults, whether these faults are initiated by natural causes such as lightning or by system malfunctions or disturbances. The use of ungrounded power circuits was commonplace at one time, and a number of distribution systems have been operated ungrounded for many years. Operating experience has shown, however, that grounded systems have better service continuity than ungrounded systems. The two main reasons for this will be discussed. While a phase to ground fault on an ungrounded system generally does not immediately result in a service interruption, the fault can cause sustained system overvoltages and, possibly, resonant conditions or restriking ground faults. These conditions will increase the likelihood of a second ground fault on one of the other phases, and subsequent power outages. Direct lightning strokes to an ungrounded power system may cause serious damage to vital equipment. The lightning current can cause severe transient overvoltages and flashovers, sometimes at several different locations on the power system. Often, these flashovers are followed by power frequency backflashes which may remain uncleared. Power system performance is not the only factor which favours grounded power systems. Ground faults on ungrounded systems can present a severe or lethal shock hazard. This is because in ungrounded systems the probability of exposure to such faults is much greater due to the longer time between the occurence of a fault and detection and removal of the faulted circuit. The importance of adequate grounding is continuously increasing with the increases in both voltage and short-circuit current levels of modem power systems. While there are still a few ungrounded distribution systems, all high voltage power systems in North America are effectively grounded. The earth is used as a return path for power frequency fault and unbalanced currents, as well as for lightning and switching surge currents. A grounding electrode performs several essential functions during normal and faulted power system conditions: 5-1 m A grounding electrode maintains the potential rise of the power system neutral below a critical value. • The grounding electrode provides an alternate path, or the only path for unbalanced faults and lightning currents. This is achieved by designing a low impedance connection between the power system neutral points and the earth. • A grounded system permits fast and sensitive detection of fault currents. By monitoring a solidly grounded system for neutral current magnitudes above a certain value, faulted circuits can be detected and isolated in the minimum amount of time. © Finally, a suitable grounding electrode is the only feasible means of maintaining the various noncurrent carrying metal structures of a power installation at safe potential levels and keeping the earth surface potential gradient in the vicinity of the electrode within tolerable values. The critical role played by a grounding electrode has not always been recognized. Many electrocutions and equipment problems may have resulted from a lack of adequate knowledge of grounding system behaviour. There are many reasons why power system grounding has been neglected as power system technology has developed. There are two factors, however, which have been the principal obstacles to the development of a thorough understanding of power system grounding. Firstly, the physical process of current conduction in the earth is three-dimensional in nature and very complex mathematically, even when point electrodes are considered. When practical power system grounds are analysed, hand calculations become impossible, unless extensive simplifications are introduced. Furthermore, it is not possible to derive empirical formulas or rules based on previously installed grounding electrodes. The reason for this lies with the complexity of the earth's structure which varies with geographical location and climatic conditions. Also the size and configuration of grounding electrodes vary within wide limits. The availability of digital computers has permitted a renewed interest in the subject of power system grounding. Numerous papers [1-22] describing computerized analytical methods for grounding analysis have been published in the last decade. Moreover, the spectacular advances in electronic technology is leading to the development of significantly improved grounding measurement equipment [23-27] which will permit a fast and accurate determination of the equivalent "electrical" earth structure and the impedance of installed grounding systems. In this chapter, the basic mathematical equations defining the behaviour of a grounding electrode injecting power frequency current into a uniform or two-layer earth structure will be derived and discussed. The analytical methods described in this chapter, although general in nature, are concerned primarily with small ground electrodes of the type usually encountered on transmission structures not equipped with continuous counterpoises. The continuous counterpoise is analyzed separately at the end of this chapter. The theory presented in this chapter is the basis for the grounding computer program GTOWER, described in Appendix C. This program was used to analyze several grounding structure configurations buried in various earth structures. It also determines the effectiveness of ground potential control conductors designed to reduce step and touch potentials around transmission structures in densely populated centers, where the probability of exposure of the public to the structure is high and should be considered. 5-2 The results of the analysis, performed using computer program GTOWER, are presented in several design charts which can be used to directly determine the performance of structure grounding systems. 5.2 TRANSMISSION LINE GROUNDING The insulation requirements of transmission lines are determined by lightning or switching surge transients and not by the power frequency voltage. In case of a severe surge, such as a direct lightning stroke on a structure, the voltage built up across a phase conductor insulator at that structure is the product of the lightning current i(t) in the structure and the structure impedance Z(t) where t represents time. The dynamic (time-varying) impedance function Z(t), is difficult to determine because of the mathematically complex structure geometry, structure footing and grounding configurations, and often, the complex nature of the soil structure. In Chapter 7, an attempt is made to derive the function Z(t). In this chapter, however, our interest is directed at the resistive part of Z(t) and, more precisely, at the structure ground resistance. Except in rare cases where soil breakdown occurs due to very high structure currents, this resistance is constant and is dependent only on the structure grounding configuration and soil characteristics. Structure ground resistance plays a major role in transmission line design. Experience and statistical data have shown that there is a definite relation between lightning outage rates and structure resistance. Consequently, the classical method for determination of transmission line outage rates caused by lightning [28,29] uses structure ground resistance as a fundamental parameter in the calculations. The importance of structure grounding has also been confirmed by measurements and theoretical analyses as discussed in Chapter 7. Analysis [28-30] shows that the potential rise of a transmission line structure struck by lightning depends mainly on the structure ground resistance, when the wavefront duration of the current is sufficiently longer than the propagation time of the surge in the structure. With very steep wavefronts, the surge response of the complete structure, including its grounding system, may influence the potential at the top of the structure, especially when the footing resistance is relatively low. In addition to the major influence it has on transmission line lightning performance, the grounding system of a transmission structure also contributes to the effective dissipation of power system ground faults. When a ground fault occurs on a transmission line equipped with overhead ground wires, a significant portion of the fault current is diverted to the structures on each side of the faulted structure. Consequently both the fault current and potential rise at the faulted structure are decreased. This has favorable effects on both transmission line performance and safety. 5.2.1 TRANSMISSION LINE STRUCTURE GROUNDING SYSTEMS The grounding system of a transmission structure consists of the following elements: • All metallic elements of the structure buried in the soil or in the concrete of the foundations. These include rebars, stub angles, guy anchors or anchor bolt cages, buried portions of structure legs, etc..• • Any supplemental grounding electrode, such as ground rods, horizontal rings, counterpoises or any combination of these ground conductors. 5-3 Although the subject is discussed in greater detail later in this chapter, it should be noted that metal structures encased in concrete usually contribute significantly to a reduction in the structure ground resistance. Often, the concrete of the foundations alone provides a suitable grounding system for the structure. This is especially true in low resistivity soils where the structure resistance may well be below 10 ohms for high voltage four-legged lattice structures, or their equivalent. The structure grounding arrangements most widely used by utilities can be grouped in two basic types (reference is made to the supplemental grounding only): • Concentrated type. • Extended or continuous type. The extended type, generaly called continuous counterpoise, consists of one, and sometimes two horizontal cylindrical conductors buried in the soil along the transmission line and connected to each structure of the line. This type of grounding system is analysed later in this chapter. The concentrated type can be subdivided into a number of categories, each characterized as an arrangement of one or more of the following basic grounding elements: a) A vertical ground rod. b) An horizontal cylindrical conductor segment. c) A ring or closed cylindrical loop. In the following sections, analytical expressions which govern the behaviour of the concentrated types of electrodes, consisting of interconnected elements of one or more categories will be derived. Although it is possible to consider a continuous counterpoise as an interconnection, along a rectilinear path, of a large number of elements from the b category, a fundamental theoretical difference, exists between the continuous counterpoise and the concentrated grounding system. While it is valid to assume that all parts of a concentrated ground are at the same potential with respect to a remote reference point, i.e., the ground conductor elements have zero self impedance, it is inaccurate to neglect the self impedance of a continuous counterpoise. For a sufficiently long counterpoise, the point of current injection is at a maximum potential while at least one extremity is at zero potential. The zero potential is due to the total voltage drop that is caused by the finite value of the self impedance of the counterpoise. This is true for dc currents as well as ac or impulse currents. In order to apply the following equations to a complete structure grounding system, some further assumptions are required. Firstly, it is assumed that small volumes of concrete buried in earth assume the resistivity of the local soil. When this assumption is not realistic, an equivalent structure can be defined whereby, the volume of concrete is replaced by an equivalent volume of the surrounding soil (see Section 5.4). Field and laboratory measurements confirm the validity of this approach, although certain restrictions must be observed when dealing with long duration currents (minutes, hours or days depending on the current magnitude) or short duration currents of very high magnitude. This restriction is due to some peculiar properties of concrete and is discussed in more detail in Section 5.4 of this chapter. The next assumption is an approximation. Since all buried metallic parts of a structure have relatively short lengths and are usually close to vertical or horizontal in orientation, they can be equated to one of the a, b or c categories identified previously, provided that any noncylindrical structures are replaced by equivalent cylindrical conductors. In addition, it is 5-4 suggested that any metallic part whose dimension is small compared to the total length of the structure grounding system be ignored, unless it represents a significant portion of the total length of the ground conductors which are buried in a low resistivity volume of soil (for instance, rods buried in a low resistivity bottom layer). 5.2.2 EQUIVALENT CYLINDRICAL CONDUCTOR The buried metal parts of a structure are generally cylindrical conductors (rebar cage, anchor bolts, etc.) or metallic beams with L, T or U shapes (lattice structure legs, foundation beams, etc.). In order to find the equivalent cylindrical conductor of these metallic structures, the well-known concept of the geometric mean radius (GMR) for n closely spaced parallel conductors [31,32] will be applied. The GMR concept, in this context, is not exact. However, it provides a convenient and accurate means of reducing a group of ground conductors to an equivalent single conductor. This technique should be used only when the distance between the conductors is small compared to their length. The GMR concept is a valid approximation because theoretical analysis shows that the effect of the ground conductor radius on structure grounding performance is minimal, even for a 4:1 change on conductor radius. Thus, even though the proposed approach is not exact, deviations from the results of a precise calculation will be insignificant for practical purposes. Figure 5.1, illustrates the principle of the proposed technique. In this figure, re is the equivalent cylindrical conductor radius, rj is the radius of the cylindrical conductor i of the group of n conductors, R is the rebar cage radius (cage center to rebar center) and djj is the distance between two conductors i and j of the group of conductors. Figure 5.1 Equivalent Cylindrical Conductor The applicable relations are as follows: Equation 5.3 is the general form of the geometric mean radius formula. The n symbol in the equation represents the cumulative product operand, i.e.; 5-5 n II a. = a* x a2------------------------x a —i s n a) Rebar Cage (assuming that rj = r): n r n e i=2 (5-1) ‘1 i b) Regular Beams (assuming that rj = r): (5-2) If the beams are regularly spaced along a straight line (spacing s), Equation 5-2 reduces to: ------------------------- 1 f:-fr r 1)l(n-2)! e 3! 2!]2 c) Irregular Shape: r (5-3) e It is also possible to determine an equivalent cylindrical conductor using the following approximate formula: re = ( 7 S)2 (5-4) Here S is the cross section of the metallic element and re is the equivalent conductor radius. This equation is only applicable to single elements and can not be used for a group of conductors such as the rebar cage. In summary, the following fundamental assumptions are made in the analysis of concentrated structure grounding systems: 1- The grounding system is an equipotential surface (infinitely conductive ground conductors). 2- Concrete resistivity is equal to the surrounding soil resistivity. If this assumption is not realistic, the concrete encased conductors can be replaced by equivalent conductors using the approximation techniques presented in Section 5.4. Therefore, this assumption is not restrictive but is made here only to avoid mixing the problem of encapsulated electrodes with the general theory of grounding electrode performance. 3- All noncylindrical elements of the real grounding system are replaced with suitable equivalent cylindrical conductors. Also, since most elements of practical transmission line grounding systems are horizontal, vertical or can easily be considered as such, the analysis assumes vertical or horizontal elements. Although there are no theoretical difficulties in dealing with oblique conductors, in the context of transmission line grounding it has been concluded that such refinements are unnecessary since they will rarely be applied in practice and will have a negligible impact on results. 5-6 5.3 LOW FREQUENCY RESPONSE OF STRUCTURE GROUNDING SYSTEMS When a dc or low frequency ac current is injected into a concentrated grounding electrode consisting of interconnected cylindrical conductors, the current flows in all conductors and is injected into earth along each conductor's surface. When the conductor diameter d is small compared to its length l and spacing D between any two conductors of the ground system, the surface current density at a given location of the conductor is, for practical purposes, uniform over the surface of the conductor (see Figure 5.2). d i—r Line Source Figure 5.2 Equivalent Linear Current Source Thus, it is possible to replace the conductor with a line source (filament). The linear density 5 of the earth leakage current in the equivalent line source is related to the surface current density a of the real cylindrical conductor according to Equation 5.5: 6 (A/m) = a(A/m2)/2irr (5"5) The linear current density is usually nonuniform over the length of the line source, being dependent on the grounding system configuration and earth structure. 8 can therefore be expressed as a function of the space coordinates x , y and z . Reference [21] provides an easy explanation of why the current distribution is generally nonuniform. 5.3.1 GROUNDING PERFORMANCE OF A LINE SOURCE In Chapter 4 of this report, the potential at an observation point, M(x,y,z), in a uniform or two-layer soil was derived for a point current source buried at the suface of the soil (Equation 4.35). This solution was obtained from Laplace's equation and appropriate boundary conditions. The solution is valid only when displacement currents in the air and in the earth are negligible. This is the case for dc or low frequency ac curents. The earth potential function describing the performance of a line source can be deduced from the expression for the potential due to a point source buried at a depth e below earth surface. This potential is obtained using an approach identical to the one used in Chapter 4 (point source at surface of soil). The results for a two-layer earth structure are (see Figure 5.3): 5-7 a) Point source in top Layer (e < h): ipi F 4>(0)+ip(e) Un(r) == -----J kn L r + Kn n=1 L ipid+K) r U12 (r) = -----------kn where and respectively. Uyi L -|‘ ^(nh)+^(nh+e) + ip(-nh)+^(-nh+e) / (5-6) r ^ ij;(0)+^(e) + J. K L i/j (nh)+^(nh+e) (5-7) are the earth potentials at a point M in the top and bottom layer b) Point Source in Bottom Layer (e > h): op ip2 U21(r) = ----- ip(0)+ip(e) EKn n=1 kn ip2 U22(r) = ----- ^(0)+^(e) r ^(-nh)+^(nh+e) - 4j(-(n-1)h)((n-1)h+e) (5-8) L £■•[ ^(nh+e) - ^((n-2)h+e) 4tt (5-9) where U21 and U22 are the earth potentials at a point M in the top and bottom layer respectively. In equations. 5-6 to 5-9 the function ^ is defined as: ^(a) =jr2 + (2a+b)2 J (5-10) where r is the distance between the point source O and the observation point M, b is the coordinate of M with respect to the vertical ow axis (see Figure 5.3) and a is the variable used to define the function W K = (p2"Pl)/(P2+Pl) Figure 5.3 Point Source Below Earth Surface Multi-Layer Earth The method used to derive the preceding equations can be extended to the n-layer earth without difficulty. The earth potential, due to a current source buried in the ith layer, at a point M in the jth layer has the general form of Equation 4-27 of Chapter 4: 5-8 •PI U. . u J0(ar)ear + 0.J0(ar)e"arda 2ir (5-11) The j3\ and B\ are determined from the boundary conditions between two consecutive layers (including air) and by the condition that the earth potential approaches zero at large distances from the point source. Of course, the solution of 2n simulataneous linear equations is required to obtain and 6\. Even for the case of a three layer earth, the analysis has become cumbersome [33-34]. Line Source A line source can be considered as an infinite number of point sources placed si de by side as shown in Figure 5.4. w Figure 5.4 POINT SOURCE SURFACE OF GROUND CONDUCTOR Infinite Number of Point sources Each point source located at a distance u from the origin of the line source injects a current i(u) into earth. When I is the total current injected into earth by the line source, the following relation holds: (5-12) I The earth potential at any point of the surface of the ground conductor must be constant since the conductor is a perfect conductor (assumption number 1). Therefore: V = o?vo ,wo)du (5-13) where V is the constant potential of the conductor. u0, v0, w0 are the coordinates of a point at the surface of the conductor. U is the potential due to a point source. U is given by one of the previous Equations 5-6 to 5-9. These equations are linear functions of the source current i(u) which is unknown but can be determined from the knowledge that the derivative of the potential function dU/du, for v2 + = a^ (a is the conductor radius), is zero (constant potential). Hence the current distribution i(u) must satisfy an integral equation for which, as yet, there is no known closed-form solution. By successive approximations a solution can be obtained. This process is already tedious for a simple horizontal conductor buried in uniform soil. However, when the grounding configuration is more complex and the 5-9 soil is nonuniform the process for obtaining a solution becomes unmanageable. In such cases, it is advisable to use other methods. The finite element technique is particularly well suited to this problem. Various algorithms have been proposed recently in the grounding literature. The most widely used method is based on the segmentation-integration process originally proposed in [4]. Another process, also described in [4], is the summation method. It involves the summation of the scalar potentials U of a finite number of point sources, which are suitably distributed over the length of the ground conductors, and the subsequent solution of a set of simultaneous linear equations to determine the unknown current distribution values i(u). This method has several major disadvantages. One of the more serious is the large number of point sources required to suitably model real-life grounding installations. The Segmentation-Integration Method This process, depicted in Figure 5.5, involves a division of the line source into a number of segments, each small enough (with respect to the overall dimensions of the grounding system) to allow the current distribution over each segment length to be considered uniform. The srrialler the length, the more accurate the results will be. There are, however, practical limitations related to computer time and memory capacity. Fortunately, experience [23] has shown that excellent accuracy is obtained when the segment length is in the order of l/10th of the maximum ground conductor length. Confirmation of this conclusion can be obtained by comparing computer results made with the same grounding system divided into a variable number of segments. Recently, model tests [21] have confirmed the validity of the segmentation process. Moreover, analysis of the test results has indicated that once a minimum number of segments is attained, further subdivision of the ground conductor becomes an unnecessary refinement. Finally, the minimum required number of segments (for a given accuracy) was found to be greatly dependent on the way the conductor was divided. In other words, there is an optimum subdivision (unequal segments) which leads to a minimum number of segments. SEGMENT j MICRO SEGMENT COORDINATES Figure 5.5 Segmentation Process Assume that the line source has been divided into a suitable number of segments. The current distribution along each segment is uniform but may vary from one segment to the next. There are a total of m segments, and each segment j injects a current Ij into earth. 5-10 Since I is the total conductor earth current, we have: m I = £l. (5-14) J-1 J The linear current density gj (A/m) of segment j of length £ j is: 6 = I /£ (5-15) Now consider an infinitely small length du on segment j of the line source (Figure 5.5). This small element will be called a microsegment and, since it can be made infinitely small (du —» 0), it can be considered as a point source. The contribution of the microsegment to the potential at a point M (uQ, v0, w0) is therefore given by one of the Equations 5-6 to 5-9. for convenience, these equations will be represented by one single general equation: U = IA [T*(a)] (5-16) where i is the point source current. A is a constant (A = PjMtt in Equation 5-6). The special character T before ^(a) means, by convention for this chapter, the appropriate sum of funtions of the type \p. is the function defined by Equation 5-10. & a is used to represent the appropriate value of the parameter defining each function ^ of the sum. The values of r and b in Equation 5-10 are determined from the following relations (refer to Figures 5.3 and 5.5). r2 = (u-Uq)2 + (v-v0)2 + (w-w0)2 (5-17) b = w0 Thus, Equation 5-10 becomes: ifKa) = j^(u-u0)2 + v-v0)2 + (w-w0)2 + (2a+w0)2J 2 The total current in the microsegment is: i . = 6.du J J and the potential contribution of the microsegment is, from Equation 5-10: dU. = 6.A J J du The potential contribution of the segment is therefore: 5-11 (5"l8) The integration of '/'(a) defined in Equation 5-18 is straightforward: {&j _u o) + \ &j - u o)2 + Vo 2 + (2a+wo) 2 Uj = 6 j ATI n (5-19) ■u 0 + Vuo2 + Vo 2 +" (2a+wo)' For example, in the case of a segment j buried in the upper layer of a two-layer earth, the potential at a point M is: oo 6 jPi Un = 4tt ■E. (£j-u0) + ^(£j-u0)2 + v02 + (2nh+w0)2' 1n - u0 + yj u02 + v02 + (2nh+w0)2 n=- (5-20) (£j-u o) + ^(£j-uo)2 + Vo2 + (2nh+2e+wo)' 1n - u0 + yj u02 + v02 + (2nh+2e+w0)2 5.3.2 ANALYSIS OF A STRUCTURE GROUNDING SYSTEM The structure grounding system of Figure 5.6 is divided into a suitable number of segments, m. Thereafter, the potential Uj at a point M, caused by a segment j, is computed using the general potential Equation 5-19 or the more explicit Equation 5-20 if the segment j and point M are in the upper layer. Since the potential is a scalar quantity, the total potential U caused by all segments (j = 1, m) of the grounding system becomes: U = m X)u. j=1 J (5-21) Note that in Equation 5-19, the coordinates uD, vQ, w0 are relative to a cartesian system attached to the segment j and with the origin at one extremity of the segment. Moreover, the u axis of the coordinate system coincides with the orientation of the segment, as shown in Figure 5.6. Therefore it is convenient to refer to only one coordinate system xyz as shown in Figure 5.6. This cartesian coordinate system is arbitrary, provided that the xy plane corresponds to the surface of the soil. The relations between the u, v, w and x, y, z coordinates are as follows: Horizontal Conductors x = xg + ucoscx - vs ina y = yg + usina +■ vcosa z = z s (5-22 (a)) + w 5-12 Vertical Conductors x = xg - w y = ys + v (5-22(b)) z = z +u s where xs> Ys ar|d zs are the coordinates of the segment origin. « is the angle between the x axis and the projection of the segment on the xy plane. Earth Surface Figure 5.6 Tower Grounding System Determination of the Segment Current Densities Equation 5-21 contains one unknown quantity, the segment current density 6j. Because the current density distribtion along the conductors of a grounding system is generally nonuniform, 5; is different from one segment j to another. As a first approximation, one may assume that the current distribution is uniform. Under this assumption the current density is: 6. = I/L (j = 1,m) where I is the total current in the grounding system (I = Sip and L is the total length of ground conductors in the grounding system (L = 2£p. This assumption may lead to significant errors, however, except for a few grounding system configurations such as ring electrodes buried in uniform soil. Therefore, in most cases where accuracy is required, it is necessary to determine the values of the current densities 5j. This is achieved by stating that the potential on the surface of all segments is constant and 5-13 equal to the potential rise V of the grounding system. Thus, if M is a point on the surface of a ground conductor, the following condition must be fulfilled: m v = ]Cmm) j=1 J Since Uj condition m E j=1 is directly proportional to dj (see Equation 5-19), we can rewrite the preceding as: 6, —Rj(M) = (5-23) V Rj/ can be defined as the mutual resistance between segment j and point M, even though there is no real physical meaning to this mutual resistance. Note that, if M is a point on segment k, and N a point on segment j, generally, £(<Rj(M) ~ £jRk(N). Rj(M) is now a known quantity and is expressed in ohm-meters Equation 5-23 must be verified for each point M on the surface of all segments, i.e., an infinite number of points in theory, and in practice, a very large number of points, say /x. Consequently, one must solve m linearly independent equations with m unknowns 8\ (i = 1, m). Since ju can be as large as desired, the system is consistent only when n is equal to m (V is assumed to be known). Now the question is which points among the many possible points should be retained. Several alternatives are possible. However, only two are of practical value. The average Potential Method (AP) In this method [35], no specific points M are selected. Rather, the average integral of the function Rj is computed for all points M along the segment k, i.e.: .fijk is the mutual resistance between segment j and segment k. In this case = The m linearly independent equations needed to determine the current densities represented by the following equation: E j=1 I £ v j k / k Mm) = 1 for k = 1 ,m Jo Xkj‘ are (5-24) J The AP method is very accurate since all points of the grounding system are indirectly considered in the equations. This method however is not retained in this study because an alternate method, which will now be discussed, is more convenient. A comparison between the AP method and the alternate method, is presented later in this section. The Center Potential Method (CP) The m points on the surface of the conductors are selected at the center of each segment j. The reader will recall that each segment j of the grounding system must have a length St j small enough to allow the current density 5j along the segment to be considered uniform and equal to the average real current density distribution 6(x, y, z). This requirement is 5-14 equivalent to saying that the values of the computed potentials at the points on the surface of the segment should remain within narrow limits around the true potential rise value, V. Since there are generally a large number of segments in a grounding system, and because the center of the segments are considered as randomly distributed over the length of the conductors, the limited number of point retained is, in fact, a statistical sample, representative of all the points. The current densities are therefore determined from the m linear equations represented by Equation 5-23 where M is replaced by M^, being the center of segment k (k = 1, m). Determining the Potential Rise and Ground Resistance So far we have assumed that the potential rise V is known, while it has, in fact, been unknown. However, by considering in Equation 5-23, the variable Xj = 6j/V instead of 5j, one can first determine the Xj using matrix inversion or other techniques [9,11-13]. V is then easily determined as follows. The total earth current in each segment is: The total grounding system current I is related to the segment currents by: m I m -E I. j=l J m yE A.£. J J Therefore (5-25) V = I Finally, the ground resistance of the grounding system is, by definition, determined from the following formula: R = V/I (5-26) Computation Methodology In summary, the following procedure is required to analyze the grounding system: a- Divide the grounding system into m small segments. b- Select a cartesian coordinate system whose xy plane is the soil surface and direct the z axis downwards toward the earth center. c- Determine the coordinates of the segment extremities (xs, ys, zs) and (xp, yp, Zp) with respect to the xyz coordinate system. Compute the angle a between the x axis and the projection of a horizontal segment (if any) on the xy plane. d- Calculate the so called mutual resistance R;^ = RjfM^) between a segment j and the center point of another segment k. An m x m matrix can be formed when j and k are varied between 1 and m. This matrix is then inverted and the 5-15 normalized current densities Xj = 6j/V are determined from the following matrix equation (equivalent to Equation 5-23): -1 1 R., X. jk mxm j m - - „ _ — The elements of the Rj^ matrix are determined using Equation 5-19 (with 5j set equal to 1) where coordinates u0, vOJ wD are related to the corresponding x0, y0, zQ coordinates according to Equations 5-22(a) or 5-22(b). The appropriate logarithmic terms to be included in Equation 5-19 (T sign) are determined from Equations 5-6 to 5-9. e- Compute the grounding system potential rise V and resistance R from Equations 5-24 and 5-26 respectively. Determine the current densities of each segment j using Equation 5-25. f- Select the coordinates (x, y, z) of the points in earth where potential calculations are required. g- Compute the potential Uj at a point (x, y, z) from a segment j. This is done using Equation 5-19 with <5j replaced by its value as determined in step e and Uqj v0, wq replaced by x0, y0, z0 according to Equations 5-22(a) or 5-22(b). The appropriate logarithmic terms to be included in Equation 5-19 (T sign) are determined from Equations 5-6 to 5-9. h- Sum all the potentials Uj according to Equation 5-21 to obtain the total potential at point (x, y, z). Chapter 10 of this volume and Part 2 of Volume 2, give several figures and design charts based on the results of computer program GTOWER. This program was designed to perform the complete grounding analysis based on the above methodology. More information on GTOWER is given in Chapter 3 and Appendix C. Comparisons Between the Average Potential and the Center Methods The main advantage of the CP method is that the same basic Equation 5-19 is used to compute the current densities and the earth potentials. In practice, this represents a significant programming simplification and, for small grounding systems such as structure grounds, it also means reduced computer time. The AP method requires Equation 5-19 to compute the earth potentials and the more complicated Equation 5-24 to determine the current densities and ground resistance. However, for large grounding systems such as those for HV substations, the inconvenience of a larger computer program is compensated for by a reduction by one half of the array size required to store the symmetric mutual resistance matrix, R^. With respect to computation accuracy of the current densities, the average potential method provides the best alternative. However, from a practical point of view, the accuracy of both methods is more than satisfactory, as illustrated in the following example. Consider a conductor of the grounding system shown in Figure 5.7(a). The true potential rise Vt along each point on the surface of the conductor is constant and therefore can be represented by a horizontal line as shown in Figure 5.7(b). 5-16 t 00 AP method CP method X—,—,—,—^ GROUND ELECTRODE 12345 ALONG CONDUCTOR SURFACE (a) Figure 5.7 (b) CURRENT DISTRIBUTION Average Potential and ELECTRODE CONFIGURATION Center Methods The AP method computes an average potential rise for each segment j of the 5 segments considered. This average value Va is indicated by the dotted line. When the potential rise at each point on the surface of the conductor is calculated using the computed average current densities, the dotted curve is obtained. The average value of the dotted curve function in each interval representing a segment is exactly Va and is represented by the dotted straight line. The CP method gives the potential Vc at the center of each segment. The value of the potential Vc (identical for all segments) is represented by a mark, When the potential rise at each point on the surface of the conductor is calculated using the computed average current densities, the broken line curve is obtained. This curve crosses the mark at each segment center location. The continuous potential curves computed using the AP and CP methods oscillate around the true potential rise value V^. The amplitude of these oscillations depends on the nature and number of segments constructed. Experience shows that when a suitable, yet practical, segmentation is realized, the maximum amplitude of the oscillations is less than 1% of the average potential value. Since the center potential values lies somewhere between the maximum and minimum values calculated according to the continuous potential method, it is concluded that the CP method also produces accurate results. There is no analytical proof to the preceding qualitative discussion, since a direct solution to the real current distribution function is not available. However, comparisons between the computation results using the AP method [10] and the CP method [5] provide an indirect confirmation of the previous discussions. Further proof is provided by the good agreement obtained between computed and measured results on scale models [21] and real-life installations [36]. In Chapter 10, some of these test comparisons and other recent tests conducted in the context of this EPRI project are described and commented on. 5-17 5.4 LOCAL HETEROGENEITIES IN THE SOIL Attention is now focused on the situation where a grounding system (or one ot its elements) is embedded in a small volume of soil having a resistivity different from the average resistivity P2 of the bulk volume of soil. This situation is quite common in practice due to either natural or man-made factors. Apart from the case where the grounding system happens to be installed in a small localized volume of different resistivity soil, a grounding system may be deliberately installed in small lakes or marshy areas in order to lower its ground resistance. Another example is where the resistivity of a small volume of soil around the ground electrode is decreased by chemical treatment or replacement of the natural soil with a low resistivity material. Finally, there are cases where metallic conductors, connected to the grounding system, are embedded in concrete which serves as the foundation of a structure (structure footings, power equipment supporting structures, etc.). Theses situations are illustrated in Figure 5.8. GROUNDING SYSTEM GROUNDING SYSTEM GROUNDING ^ SYSTEM — Pi RANDOM HETEROGENEITY DELIBERATE INSTALLATION IN LAKE OR MARSHY AREA CHEMICAL TREATMENT OR SOIL REPLACEMENT METALLIC CONCRETE STRUCTURE FOUNDATIONS (b) ARTIE I CAL HETEROGENEITIES (a) NATURAL HETEROGENEITIES Figure 5.8 Local Soil Heterogeneities A rigorous analysis of the performance of a grounding system totally or partially embedded in a relatively small volume of soil whose resistivity p]_ is different from the bulk earth resistivity P2, is only possible for simple ground systems and certain configurations of soil heterogeneity. Fortunately, when an exact solution is not feasible, it is possible to gather very useful information through judicious approximations. In order to make the appropriate approximations, it is necessary to analyze the following fundamental problems. It is assumed unless stated otherwise, that soil resistivity is independent of the injected current and its duration. 5.4.1 HEMISPHERICAL ELECTRODE EMBEDDED IN A HEMISPHERICAL SHELL This problem, depicted in Figure 5.9, was resolved in Chapter 4, Section 4.3.6, for a point source. The potentials in medium 1 and U2 in medium 2 are given by Equations 4-43 and 4-44 respectively. These equations are also valid for hemispherical electrodes. In this case, the ground resistance R is obtained as follows: IR = ll(r ->-«>) - U(r = a)=U2(r->00) - Ui (r = a) From Equations 4-43 and 4-44, the following is obtained: R + P2 " Pi (5-27) A 5-18 MEDIUM 2 Figure 5.9 MEDIUM 1 Hemispherical Configurations The resistance of a hemispherical electrode in uniform soil of resistivity P2 is given as: 1 p2 R2 = — ( — ) 2tt a Thus, the effect of the heterogeneity can be expressed as the ratio of R/R2. This ratio a is less than unity when medium 1 is more conductive than medium 2. We then have: a a 1 + P2 - (— A pi Pi 0 (5-28) — p2 Figure 5.10 gives the value of a as a function of the ratio of the resistivities, for several practical values of a/A. This figure shows that in order to reduce the resistance by half (a = 0.5), it is necessary to reduce the resistivity of a hemispherical volume of soil of about 2.5 times the electrode radius by l/10th. Figure 5.10 can be used to determine the equivalent radius ae of the electrode which, when buried in a uniform soil of resistivity P2, will have the same resistance as the electrode embedded in the hemispherical heterogeneity of resistivity Pl. This equivalent radius is: a e = a/a (5-29) Of course, the earth potentials around the equivalent electrode will be different from the potentials which exist around the original electrode, particularly at points close to the boundary between medium 1 and medium 2. However, when the volume of the heterogeneity is small with respect to the grounding system dimension, this concept of equivalent radius can be used to approximate the effect of the heterogeneity. To illustrate this, let us consider the transmission line tower shown in Figure 5.11. In this figure, we have assumed that the concrete in the foundations has an hemispherical shape. Other shapes will be considered later. We also assume that the resistivity in the concrete is significantly different from that of the remainder of earth, although this is seldom the case in practice for small volume of concrete. Firstly, the metallic elements of the foundations are replaced by an equivalent hemispherical electrode. This is accomplished using the following approximations: a- The metallic elements of one leg are considered as buried in a uniform soil of resistivity p^. b- The resistance R of this leg is computed. c- The equivalent hemispherical radius Pl/27rR. 5-19 a is deduced from the relation a = RESISTANCE RATIO (p.u.) Next, the resistance ratio a of the structure leg foundations (metallic elements and concrete) is determined from Figure 5.10. Equation 5-29 gives the required equivalent radius ae- HEMt SPHERICAL HETEROGENEITY EQUIVALENT RADIUS RESISTANCE RATIO a/A=0 EQUIVALENT RADIUS = a/a Figure 5.10 Effects of a Hemispherical Heterogeneity STRUCTURE FOUNDATION Figure 5.11 Tower Foundations .2 5.4.2 VERTICAL ROD EMBEDDED IN A CYLINDRICAL SHELL Hemispherical electrodes are rarely used in practice. Moreover, very few grounding systems can be approximated by a hemispherical electrode. Most grounding systems consist of cylindrical conductors with a radius significantly smaller than the conductor length or spacing between conductors. In most cases, the buried elements of a metallic structure are connected to the grounding system and therefore are part of it. These metallic elements are generally linear and have a small cross-section relative to their length. Often, several vertical elements of the structure are embedded in a small volume of concrete. These comments suggest that the case of a vertical rod embedded in a small cylindrical volume of a material with a resistivity different from that of the local soil, is of considerable interest. This problem has been thoroughly investigated in [37]. However, the proposed method did not consider the distortion of the equipotential surfaces due to the lower extremity of the rod. In the following analysis this distortion effect is approximated using the cylinder-hemisphere concept originally proposed in reference [38]. This concept is illustrated in Figure 5.12. The thickness of the material enclosing the vertical conductor is assumed to be small compared to the conductor length. Thus, the earth current flowing out from the conductor surface can be assumed to flow radially (perpendicular to the conductor surface) within medium 1, since it is assumed that the rod and medium 1 surfaces are concentric as shown in Figure 5.12. Thus, both surfaces, as well as any intermediate concentric surfaces, are equipotential. The resistance between any pair of these surfaces can then be computed using the conventional relation between resistance and resistivity. Figure 5.12 Small Cylindrical Heterogenous Material Assuming that the lower extremity of the vertical rod is hemispherical, the resistance dRj between two infinitely closely spaced equipotential surfaces (located at r and r+dr respectively) is given by the relation: length dr dRi = resistivity ------------------------- = Pi -------------------cross-section 2Trr£ + 2 tit2 5-21 The total resistance between the conductor surface and the boundary surface between medium 1 and medium 2 is therefore: A{a + H) Pi = ------ 1 n Ri (5-30) a (A + £) IttI The difference between this resistance and the resistance R2 obtained using Equation 5-30 with a different resistivity P2 (rod in uniform soil), is given as: (Pi - P2) A (a + l) AR = Rj - R2 = -------------- In ------------2Tr£ a(A + £) (5-31) The ground resistance of a vertical rod with a radius a given with excellent accuracy by the following equation: much larger than its length £ is p2 21 R = ------ In — Ziri a (5-32) This equation assumes a uniform soil of resistivity P2. The actual resistance of the rod embedded in the material of resistivity P^ which in turn is surrounded by the soil with a resistivity P2 is therefore: 2£ P2 R 1n = R + AR = 2tt£ + (pi “ P2) - -------------- 1 n (1 + £/a) 2tt£ _(1 + £/A). (5-33) Equation 5-33 can then be used to determine the equivalent radius of the vertical rod which, when buried in a uniform soil of resistivity P2, will have a ground resistance equal to Rt. This is achieved by equating (5-33) with (5-32) where a is replaced with ae. This leads to: 1 + £/a a e (1”Pl/p2) (5-34) a 1 + £/A Inspection of Equation 5-34 reveals that when P]^ < P2, the equivalent radius is larger than the actual vertical rod radius. The reverse is also true. This is as expected, since the resistance decreases when the radius increases. Figure 5.13 shows plots of the ratio, /? = ae/a, as a function of the resistivity ratio P2/P1 for several practical values of the radius ratio a/A. The results of this figure are based on a rod length to radius ratio of 20 (£/a = 20). For a group of rods, the equivalent radius concept must be used prior to establishing the ratio j8. It is important to note that a number of assumptions and approximations have been made in order to determine the equivalent rod radius as determined from Equation 5-34 or from Figure 5.13. Careful examination of a problem, together with sound engineering judgment, is the only reliable method of determining whether or not the preceding approximate techniques are applicable. For instance, if the value Pj of medium 1 is only slightly different from P 2, it is certainly more expedient in ignoring the disturbing medium, with a minimal loss in accuracy. 5-22 a/A = a/A = 1 EQUIVALENT RADIUS ae RADIUS RATIO RESISTIVITY RATIO p2/p RESISTIVITY RATIO (p.u.) Figure 5.13 Effects of a Cylindrical Heterogeneity 5.4.3 VERTICAL ROD ENCASED IN AN ELLIPTIC SHELL The cylinder-hemisphere equipotential surface concept, presented in the previous section, can readily be demonstrated to be a valid approximation. At large distances from the rod, the equipotential surfaces are practically hemispheres. At short distances, these surfaces follow the assumed cylindrical-hemispherical shape of the rod. At intermediate distances, however, the shape of the equipotential surfaces around a cylindrical rod is, in fact, more complicated. For this reason the analysis in Section 5.4.2 was restricted to small volumes of material in medium 1. It is believed, however, that the method is still reasonably accurate when moderate volumes are considered. As will be shown, in a uniform and isotropic soil, the equipotential surfaces around a cylindrical, vertical conductor can be represented with a high degree of accuracy as semi-ellipsoids of revolution. This property has been used in [39] to investigate the effectiveness of conventional chemical and mechanical treatment of the local soil around grounding electrodes in order to lower their ground resistances. A method based on the elliptic equipotential surface concept is not only more accurate than the previous one, but is valid for any volume of medium 1, provided that the shape of this volume corresponds, or can be approximated by, one of the equipotential ellipsoids. This case is depicted in Figure 5.14. 5-23 Figure 5.14 Elliptic Shell The equation of an ellipsoid with center at the origin of a rectangular coordinate system xyz with semi-axes lengths C, D, B is (see Figure 5.15): When C = B, i.e., the ellipsoid is an ellipsoid of revolution with respect to the axis y. In this case we have: x2/D2 + t2/B2 = 1 (5-35) t (5-36) where y2 + z2 ' We also have the following relation between the major and minor semi-axes D and B and the distance from center o to focus F or F': D2 = B2 + l2 Figure 5.15 (5-37) Ellipsoid 5-2b As has been shown, the earth equipotential surfaces around the vertical rod shown in Figure 5.14, are semi-ellipsoids of revolution. More precisely, the lower focus of the equipotential surface is exactly at the lower extremity of the rod. Consequently, for each equipotential surface defined by its minor semi-axis B (distance from the rod to the equipotential surface), the major semi-axis D (depth of the equipotential surface) is related to B, and to the rod length, according to Equation 5-37. The preceding statement will now be proved. The potential caused by a segment of conductor at any point (u, v, w) in the first layer of a two-layer earth, is given by Equation 5-20 where u0, v0, wQ are replaced by u, v, w. When the conductor consists of the vertical rod of Figure 5.14 and its image with respect to the earth surface plane, (assuming that the conductor is buried in an infinite medium of uniform resistivity, as in the case of Figure 5.15), Equation 5-20 simplifies to: pi (L-u) + ^(L-u)2 + t2 U = ----- In -----------47TL ------ -u + \u2 (5"38) + t2 where, I is the total conductor current, L its length (L = 2 £), resistivity and t is given by Equation 5-36. P is the medium The equipotential surfaces around the conductor are defined as U(u,y) = constant, which is equivalent to: (L-u) +V(L-u)2 + t2' Y = ------------ r -u +\u (5-39) ------+ t This can be simplified to: (Y-1)2 yl2 u(u-L) + ---------- t2 = ---------by (y-1)2 Since, u = x + H and L = 2£, this last equation becomes: Ky + D (y - D iiyfy' (y - 1) Equation 5-40 has the same form as Equation 5-35, and therefore, we can conclude that: • The equipotential surfaces are ellipsoids of revolution. • The center of the ellipsoid is at the center of the conductor consisting of the rod and its image. Since the problem is symmetrical with respect to the earth plane, this center is also at the upper extremity of a ground rod. • The major and minor semi-axes of the ellipsoid, D and B, are related to the constant y according to the following equations: nyfT B = ----------(y-1) l{y + 1) D = ---------------(y- 1) 5-25 (5-41) from which: D2 = B2 + l2 therefore, from Equation 5-37 we can state that: • The lower extremity of the rod is a focus of the ellipsoid. Consider now the vertical rod shown in Figure 5.14. The resistance of the soil between the conductor surface and the boundary surface between medium 1 and medium 2 is: Uc " Ub cb where Uc is the conductor surface potential and Uj-, is the boundary surface potential. These potentials are determined from Equation 5-20, i.e., for a uniform soil and a ground rod driven to a depth £ in the soil: (£-u) Pil U = ----- In +yj(£-u)2 -u + -^u2 + v2 + w2 2tt& potential of the rod (u = o ; £ Pil = ----- In c 2ir£ U + v2 + w2 +yjn2 v^ + = + a2 (5“42(a)) a and since a « £: U 2£ Pil = -----c 2ir£ (5-42(b)) In La _ The potential at the boundary surface (u = o ; Pil Ub = v^ + w^ = B^) is: £ +VF + B" 1n 27r£ which, based on Equation 5-37 simplifies to: Pil Ub = £ + D In (5-^3) 2ir£ therefore: 2£' Pi cb 1n '£ + D (5-44) - 1n 27r£ B The resistance between the boundary surface and remote ground (at zero potential) is: £ + D Pzi boo 1n 2tt£ (5-45) . B 5-26 Thus, the total ground resistance of the encapsulated ground rod is: R Pi ------ 1 n 27t£ t 2£ — a (P2-P1) + ------------- In 2ir£ -1 (5-46) B J The equivalent radius of the rod buried in a uniform soil of resistivity P2 has the same resistance defined as: P2 R = ----- In 27r£ 2£ (5-47) a from this: a I" 2£/a = a I ------------e I (£+D)/B (I-P1/P2) The radius ratio /8 = ae/a is plotted in Figure 5.16 as a function of P 2/ P1, for several values of a/B, and for the case where £/a = 20. Equation 5.47 is valid provided that a « £ . When the rod radius can not be neglected with respect to its length, the equivalent radius formula should be derived using Equation 5.42(a) instead or 5.42(b). 5.4.4 COMPARISON BETWEEN THE EQUIPOTENTIAL SURFACE METHODS In order to permit meaningful comparisons between the results of Figures 5.13 and 5.16, the volume of medium 1 should be the same in each case, i.e.: 4 4 'irA2£ + — tiA3 = — ttET 4B2 + £2 6 6 This simplifies to the cubic equation: c2y3-(£2/a2)y -1=0 (5~48) where y = (a/b)2 A2 / 3£ A c = 7U+r Equation 5-48 has only one positive real root. This provides the approppriate minor axis B for each value A of the radius of the cylindrical surface considered in Figure 5.16. Figure 5.17 shows the radius ratio (/3 = ae/a) computed using the cylindrical (dotted curves) and ellipsoid (solid curves) methods. These curves were derived using the equivalent volume concept. As expected, there are significant differences between the results. 5-27 ft/a = 20 a/B = a/B = 1 EQUIVALENT RADIUS ae RADIUS RATIO RESISTIVITY RATIO p2/p RESISTIVITY RATIO (p.u.) Figure 5.16 Effects of an Elliptic Heterogeneity •0.155 - a/A •0.722 a/A = a/A = a/B = a/A - 1 EQUIVALENT RADIUS ae RADIUS RATIO RESISTIVITY RATIO p2/pi RESISTIVITY RATIO (p.u.) Figure 5.17 Equal Volume Base 5-28 The ellipsoid method is, in theory, more accurate than the cylindrical method. However, in practice, the shape of the excavation containing medium 1 (concrete, soil replacement, etc.) is generally rectangular or cylindrical. Consequently, the real equipotential surfaces in the vicinity of the rod are distorted somewhat, and their shape is probably intermediate between a cylindrical-spherical surface and an ellipsoid. Therefore, the radius ratio /3 should be selected using the curve which represents the average of the shaded area. The hemispherical approach is probably useful when chemical treatment of the soil around the electrode is used to decrease the local soil resistivity. It is clear from the previous discussion, that in practice, no method can be considered as superior to the other. Depending on the problem, one method may be more appropriate than the other. In many cases, the three concepts can be advantageously applied to determine a suitable equivalent radius. Briefly stated, the three methods described in the preceding sections should be considered as useful, but approximate, techniques, and should be applied with good engineering judgment. Additional curves based on different values of the ratio £/a are given in Volume 2. 5.4.5 EXTENSIONS AND LIMITATIONS OF THE EQUIPOTENTIAL SURFACE METHODS The hemispherical, cylindrical, and ellipsoid methods are all based on a well-known technique from electrostatic theory. This technique is derived from the principle which states that any equipotential surface (U = constant) can be replaced with a thin metallic surface maintained at the potential U, without altering the problem. Thus, when the equipotential surface around a grounding system is known, the resistance of the grounding system may be obtained. The resistance is equal to the resistance between added to the resistance of the grounding conductor surface Sc and the boundary surface the equipotential boundary surface (assumed to be a thin metallic sheet) buried alone in the natural soil. For application of this electrostatic principle, the equipotential surface boundary between the material, in which the grounding system is embedded, and the remainder of natural soil, must be assumed to be equipotential. This procedure is illustrated in Figure 5.18. ELECTRODE SURFACE Sc ^ ACTUAL 1 BOUNDARY SURFACE S REMOTE HEMISPHERICAL EQUIPOTENTIAL SURFACE APPROXIMATE BOUNDARY SURFACE (a) ACTUAL PROBLEM (b) Figure 5.18 EQUIVALENT PROBLEM The Equipotential Metallic Sheet Principle 5-29 It should be noted that, although the equivalent problem depicted in Figure 5.18(a) cannot provide the earth potential values in medium 1, the earth potentials in medium 2 remain identical to the real case. Thus, the problem reduces to determining the equipotential surfaces around the grounding system. This is a complex problem in uniform or multi-layer earth structures and, in most cases is impossible to resolve when localized earth heterogeneities are present. Our effort is not vain, however, if we restrict ourselves to heterogeneities enclosing the ground system conductors, particularly, as in the case of transmission line grounding, where concrete and other backfill materials are used. In such instances, suitable approximations, together with the analytical techniques presented in this chapter, yield reasonably accurate solutions. We will now discuss the step-by-step procedure required to achieve this goal. The type of problem to be analyzed is shown schematically in Figure 5.19. The problem is limited by the following assumptions: a) Except for the localized heterogeneities, the earth is a uniform or two-layer structure. Multi-layer earth can be handled in a similar fashion, but the equations are considerably more cumbersome and will require considerably more computing time. b) The earth potentials in the volume delimited by the heterogeneities are not of interest in these calculations, but rather will be determined in a separate step (step 8). Figure 5.19 Illustration of a Typical Problem The procedure is as follows: Step 1- Analyze the problem as if there are no localized heterogeneities, using the analytical methods provided in this chapter. Let Rc be the computed ground resistance. Step 2- Determine the equipotential surfaces around the grounding system up to the equipotential surface which, approximately (on an average volume basis, for example), encloses all the heterogeneities. Step 3- Replace all materials which exist between the ground conductors surface Sc and the equipotential surface Sfo, with an equivalent average resistivity pa based on the respective resistivity and volume of each single material removed. 5-30 Step 4- Knowing that the earth current stream lines are orthogonal to the equipotential surfaces, compute the total resistance Rcb between the grounding system surface and surface S{-), by adding the elementary resistances calculated between two closely spaced equipotential surfaces located at r and r + Ar. In this step, it may be necessary to subdivide the volume between these closely-spaced surfaces into smaller regular volumes. Step 5- Replace the equipotential surface with a metallic cage consisting of cylindrical metallic conductors. This cage should resemble, as much as possible, the shape of the equipotential surface S^. Step 6- Compute the ground resistance Rboo of this cage when buried in the earth free of the localized heterogeneities. Let V^oo be the earth potentials at any point outside this cage. Step 7- The approximate resistance of the grounding system in the real earth structure is then: R,n = R cb, + R,b°° The actual potential rise of the grounding system is therefore: uh = y where I is the grounding system current. The actual potential value of the equipotential surface Sj-, is: U:=U t> c - R cb l = R cb J and the earth potentials at any point outside this equipotential surface are: “b - W(IRbJ = V(lW.-'> Although this may appear to be a simple procedure, considerable computational effort is required to determine the equipotential surfaces. Thus, unless the maximum possible accuracy is required, the design approach should commence with the determination of an equivalent model, based on the spherical, cylindrical or ellipsoid approximations. When these methods are not applicable, the real problem should be simplified to a problem requiring a minimum of computation in each step. Fortunately, transmission line grounding problems are, in practice, amenable to such approximations. As mentioned at the beginning of this section, one further step may be required in some cases. This additional step is required to determine the earth potentials within the volume delimited by the equipotential surface Sj-,. This is a very approximate procedure when dealing with several isolated heterogeneities of dissimilar materials as illustrated in Figure 5.19. The earth potential at a point M within this volume is approximated as follows, however: Step 8- Determine the equipotential surface Sm which traverses point M. Sm is selected from the surfaces already determined in step 1. Next, as in step 4, compute the resistance Rcm between the grounding system surface and surface Sm (based on average resistivity P a). The potential of surface Sm is: This is also the earth potential at point M. 5-31 5.4.6 CONCRETE ENCASED GROUNDING ELEMENTS. Most of the commonly used metallic transmission line structures are made of steel and are of the four-legged lattice type (towers) or of the single shaft tubular type (poles). Typical foundation arrangements for these structures are shown in Figure 5.20. The tower foundation consists of a concrete pier where one leg of the tower is embedded. Depending on the mechanical requirements, there may be reinforcing steel bars in the concrete. The steel pole foundation usually consists of a cylindrical anchor bolt cage embedded in a concrete pier and surrounded by a number of reinforcing bars, also embedded in the concrete. Figure 5.20 Typical Concrete Foundations There is considerable controversy on the current conduction properties of concrete [41-44], The facts that concrete consists of a variable mixture of materials of different grain sizes and that concrete may be buried in various types of soil, might explain why contradictory results have been reported in the literature. There is ample experimental evidence [40,41] which shows severe damage to concrete as a result of sustained or short duration alternating currents flowing from concrete encased conductors. However, other experimental tests and considerable operating experience [37,42-44] shows, equally well, that concrete encased ground conductors have very substantial ground current capability and, therefore, are as effective or sometimes superior to equivalent conductors in direct contact with soil. Both results are credible and differ only because of the duration and magnitude of test currents and soil conditions. The moisture content is a particularly important factor. There are other factors which influence the conduction process to a lesser degree. As explained in Chapter 4, current conduction in earth is primarily a function of the electrolytic content of soil, i.e., presence of water and salts, acids, etc. Similarly, concrete resistivity is a function of its moisture content. However, due to its inherent alkaline composition, hygroscopic nature and density, it is less sensitive to the content of salts or acids in the soil moisture and thereby, is more dependent on its inherent water content. Since concrete in the earth tends to draw moisture from the local soil, concrete resistivity is comparable to or lower than the local soil resistivity, provided that the volume of the concrete is moderate enough to permit water infiltration through the whole volume. When wet, concrete resistivity values range from about 30 fi-m to approximately 300 fi-m. When dry, concrete is a very poor conductor with resistivity values ranging from a few kfl-m to more than 10 kf2-m. 5-32 Conduction Current Process in Concrete Recent tests conducted in Poland concerning the effects of lightning and alternating currents on reinforced concrete foundations [44] have identified three different mechanisms of electric conduction through the concrete: (1) electrolytic, (2) spark discharges, (3) arc discharges. The results of the tests show that concrete foundations are able to repeatedly conduct very high lightning currents without noticeable damage, but may be severely damaged during high momentary short-circuit currents or sustained low magnitude fault currents [40]. In general, it appears that such damage is possible only when a certain critical energy distribution in the concrete is attained during the current conduction process. When the current density and the electric field gradient within the concrete are low, the conduction process is purely of the electrolytic type and the ratio of voltage to current is constant, i.e.: v(t) R (t) = ap (t) = ------- = constant i(t) where, v(t) and i(t) are the applied voltage and resulting current, R(t) is the dynamic resistance, p(t) is the apparent resistivity, a is a constant depending on the means of injecting test current into the concrete, and t is the time. When the test voltage and current are progressively increased, the energy dissipated in the concrete starts to evaporate the water contained in the voids and microcracks of the concrete. As a result, the potential gradients at the location of the evaporation increase significantly (5-8 kV/cm) and cause spark discharges which quickly propagate within the concrete volume. Concurrently, a significant reduction in the electric field follows the occurence of spark discharges. Therefore, the dynamic resistances and apparent resistivity of the concrete decrease substantially. If the test current is further increased, the least resistance spark channel will quickly transform into an arc channel where most of the current will flow. This arc channel may cause severe damage to the concrete foundation. Figure 5.21 shows typical dynamic resistance curves, R(t), obtained during laboratory experiments [44]. These curves clearly show the electrolytic conduction zone, a transition zone where the dynamic resistance decreases rapidly, and the spark discharge zone, where concrete behaves as if its resistivity is considerably reduced from the static value. 12 14 16 18 CURRENT IN KA Figure 5.21 Dynamic Resistance of Concrete (Redrawn from [44]) 5-33 Each curve shown in Figure 5.21 corresponds to a test current with a different peak value. During these short duration current tests (which emulate lightning currents), no visible damage to the concrete was observed. The concrete surface current density reached about 3 A/cm2 (20 A/in2). Other tests show that a current density from 5 A/cm2 to 10 A/cm2 (30-60 A/in^) is required to cause visible damage. When the current density exceeds 10 A/cm2, arcs initiate fusing of the concrete.The result!no pressure build up due to excessive heat in the voids causes explosions which are directed towards the outside of the foundation. These result in the formation of craters which may be up to 4 to 6 cm in diameter (1.5-2.5 in) and 2 to 3 cm in depth (about 1 in) if the current concentrates in a single arc channel. All of these tests indicate that no damage to the concrete foundations should be expected when the surface current density is below 5 A/cm^, that some damage will occur at values ranging from 10 to 15 A/cm^ and, finally, severe damage can be expected when the current density exceeds 15 A/cm^. It should be noted that the current density is not the only criteria. The duration of the current is also critical. During impulse currents, the safe limits are significantly higher, while under adverse conditions, substantially lower sustained currents may cause considerable damage. The limiting current density values given in this paragraph generally apply to power frequency short-circuit currents from 0.1 to 0.5 seconds duration. Practical Situations The results reported in the preceding section are based on laboratory measurements which were not intended to duplicate conditions existing in actual installations. In the following, the current densities which are likely to exist in the concrete foundations of a transmission line structure during fault conditions are examined. Impulse Current Lightning currents rarely exceed 200 kA. Based on the dimensions of modern high voltage structure foundations, the concrete current density will not exceed 20 A/cm2 if 0ne foundation is considered and 5 A/cm2 when the four leg foundations are involved. Considering the facts that high magnitude lightning currents are of extremely short duration, that concrete can withstand a much higher voltage under impulse currents and that the transmission structure usually has supplemental grounding elements and is often connected to a ground wire, both of which can divert significant portions of the current, it can be concluded that concrete damage due to lightning currents is unlikely except for relatively small isolated structures. This conclusion is supported by many years of operating experience over thousand of miles of transmission lines. Momentary Currents Short-circuit currents in the order of 100 kA are very possible now or in the near future. Their duration is limited to few cycles by modern high-speed relays, however. Based on the large dimensions of foundations required to support such high voltage structures, a concrete surface current density of 2 to 4 A/cm2 is an upper limit, even when supplemental grounding elements and overhead ground wires are disregarded. A more typical value of 1 A/cm^ is significantly below the safe limits quoted previously. Consequently, no damage should result from short-circuit currents flowing in the concrete foundations of high voltage lines. This is also confirmed by operating experience. There is a situation, however, which may lead to a considerable increase in the concrete surface current density. This adverse situation occurs when metallic structures (such as pipelines), not connected to the transmission structure, are buried close to one of the faulted structure foundations. In this case, most of the fault current concentrates in a narrow channel along the shortest path from the foundation to the metallic structure. Surface current densitites may reach hundreds of A/cm2. Sustained Currents Sustained fault currents rarely occur on high voltage transmission lines. This type of fault is likely to occur on distribution lines not equipped with overhead ground wires and when the fault occurs at a large distance from the distribution transformer or when the faulted structure is poorly grounded. Under these very unfavorable circumstances, the continuous flow of current through the structure may cause severe damage to transmission line tubular steel (or reinforced concrete) poles with relatively small concrete foundations. In some documented cases, severe arcing has occured and resulted in the destruction of concrete poles [40,41]. The arcing phenomena referred to is shown in Figure 5.22 for reference. I- TRANSITION ZONE REGION OF HIGH TEMPERATURES REGION OF SPARK DISCHARGES REGION OF ARC DISCHARGES Figure 5.22 Arcing Process in a Concrete Distribution Pole (Redrawn from [40]) As a general conclusion concerning the behaviour of the concrete foundations of transmission line structures conducting fault currents, it can be safely stated that no damage is likely to occur under any type of fault which causes current circulation in the foundations. Therefore, no special precaution is required to separate the structural metallic elements or rebars in the foundations from the supplemental grounding elements. As a guide to allowable current densities, an upper limit (conservative) of 5 A/cm2 can be used for current durations in the order of 0.1 second. It is suggested that below 0.1 second the product of the current density in A/cm^ and the square root of the time in seconds, should be kept below 2 to remain within the same conservative test result limits. 5.5 CONTINUOUS COUNTERPOISES In the foregoing sections, the ground conductors have been assumed to be of such short length that the voltage gradient along their length can be neglected. For lengthy grounded conductors, it is necessary to consider the longitudinal voltage drop along the conductor. Cable sheaths, railroad tracks, pipelines and transmission line continuous counterpoises are typical examples of long metallic conductors totally or partially in direct contact with soil. The current flowing in such conductors will induce a voltage drop due to the internal impedance of the conductor. 5-35 The continuous counterpoise consists of one or two conductors buried continuously under the transmission line or along certain sections of the line. The conductors are interconnected to the overhead ground wire and ground system (if any) at each transmission line supporting structure. 5.5.1 PERFORMANCE OF AN ISOLATED CONTINUOUS COUNTERPOISE. Figure 5.23 shows a continuous counterpoise of total length & = buried in a uniform soil at a depth e below the earth's surface. A low frequency alternating current 1° = Ii + l£ is injected in the counterpoise at a point A corresponding to u = o and w = e with respect to the uvw coordinate system, where the u axis is parallel to the counterpoise and the plane defined by the u and w axes is the earth's surface. The return current electrode is assumed to be located at a large distance from the counterpoise (remote earth) so as not to interfere with its performance. Finally, it is also assumed that there are no other buried metallic structures (isolated counterpoise). U° 0 Mr _______ ---------- “T-TPFT” ; ! u I ^! a ! ♦ W REMOTE (POTENTIAL = 0 V) Figure 5.23 Isolated Continuous Counterpoise The current entering the left and right side of the counterpoise at the injection point A(u=o, w=e) are I^ and if respectively. Let I(u) and U(u) be the current and earth potential of the counterpoise at distance u . Figure 5.24 shows an enlargement of an infinitesimal element of the counterpoise between distance u and u + du. I(u+du) I(u) — *- U(u) —njuL—'wows'— c = ■[]6 U(u+du) u+du Figure 5.24 Infinitesimal Counterpoise Element 5-36 The application of Kirchoff's laws to this elementary circuit leads to the following expressions for the current and potential of the conterpoise at distance u + du: dI(u) = l(u + du) - I(u) = - (G + jojC)duU(u) (5-^9) dU(u) = U(u + du) - U(u) = - (r + ja)A)dul(u) (5-50) where r, X, G and C are the electrical parameters of the counterpoise, i.e.: r is the unit length resistance (ohms) X is the unit length inductance (henrys) G is the unit length shunt conductance (to ground),(mhos) C is the unit length shunt capacitance (to ground),(farads) The current I, or the potential U, can be eliminated from the pair of simultaneous Eguations 5-49 by differentiating with respect to u. When this procedure is applied to the counterpoise current, the classical wave equation is obtained: d2I --------- a2I = 0 (5-51) du where z = r + ja)A Y = G + JojC a is the complex propagation constant. The general solutions of the differential Equation 5-51, assuming that z and Y are independent of u , are: T au , -au 11 = aie + b1e ; , . when u < 0 , au -au 12 = 826 + b2e ; . . „ when u > 0 Two similar, but different, expressions are required for the counterpoise (longitudinal) current, since this current is not defined at u = 0. It is defined, however, at u = ± e, where e is an infinitely small positive number: 11 = Ii ; atu=e 12 ; atu = + e ~ 12 The constants are determined from the boundary conditions: Ii = 0 9 when u = -Al 12 = 0 9 when u = £2 II = I? when u = 0" I2 = 1° when u = 0+ and Ui(u=-e) = U2(u=+e) = u' when e 5-37 0 This last boundary condition states that unlike the current I, the counterpoise potential U is continuous around u = 0. From Equation 5-50, this condition is equivalent to stating that: / 1 dlA / 1 dl2\ \ V du / u=+e = U° \ Y du / u=-e when e 0 after some elementary algebraic transformations, the following solutions are obtained: u < 0 cosh(aiz) ii = -r , i nh|a(Jli + u)J (5-52) i i nh(a£) u > 0 cosh(a£i) inhja(£2 " u)J I2 = I1 (5-53) s i nh (ail) The counterpoise potential is now easily determined from Equation 5-49. The following solutions are obtained: u < 0 cosh(a£2) r -| Ui = I°g --------------- cosh a(£j + u)J sinh(a£) (5-54) u > 0 cosh(a£i) U2 = I 0 ( .... T cosh j^a(£2 - u)] (5-55) a(iz sinh(a£) where = 'Jz/Y The potential rise of the counterpoise at the current injection point (u = 0) is therefore: cosh(a£i) cosh(a£2) U° = I°f sinh(a£) and the ground impedance of the counterpoise as seen from this current injection point is: cosh(a£i) cosh(a£2) Z = U°/I0 = s i nh |^a(£i + £2)J Specific Cases and Practical Considerations When the injection point is at one extremity of the counterpoise (£j^ = 0 ; ground impedance is: Z = 8 coth (a£) = ■^z/y'coth zY£ £2 £), this (5-56) 5-38 When the injection point is at the center of the counterpoise, (Equation 5-55 reduces to: 3 Z = — coth(a£/2) V (5-57) z'/Y coth 2 where z' = z/k A comparison of Equation 5-56 and 5-57 shows that energization of the counterpoise from its center is equivalent to reducing the self impedance of the counterpoise by four times when energized from one of its extremities. Thus, the counterpoise is applied more effectively when energized from its center. When the product VzY i is large, the hyperbolic cotangent is approximately equal to unity. The counterpoise ground impedance is then: Z =^/z/y‘ (5-58) This is the case for very long counterpoises or when a = VzY is large. At power frequencies, the counterpoise capacitance C can be neglected compared to its conductance G . Therefore, Y ~ G = l/Rj, where Rj is equal to the counterpoise distributed ground resistance in ohms/meter. Hence, the ground impedance given by 5-58 becomes: z The distributed ground resistance, Rj (per unit length), of a counterpoise can be calculated using the general equations developed in Section 5.3, for a unit length of conductor. When the counterpoise is buried close to the surface of a uniform earth, the ground resistance is approximately: P 21 R = — In _ Ttl ■y 2ae (5-59(a)) where l is the counterpoise length, a is the radius and e is the depth of burial. Therefore, the counterpoise distributed ground resistance per meter of conductor is: P R d 21 ' (5-59(b)) R£ TT yfae Figure 5.25 gives the modulus of the ground impedance of a counterpoise buried in a uniform soil with respect to its length £, for several values of soil resistivity p and for injection points at one extremity and the center of the counterpoise. Figure 5.26 shows the effect of frequency on the ground impedance of a counterpoise of length £, buried in uniform soil of 100 fl-m resistivity and for an injection point at one extremity. Additional ground impedance curves are also provided in Volume 2 of this report. The characteristics of the counterpoise are as follows: Conductor: copperweld conductor Radius a: 0.5 cm (0.2 inch) Depth of burial e: 0.5 m (20 inches) 5-39 Capacitance C: 0.03 MF/km (0.03 MF/mi) Resistance r: see next paragraph Inductance X: see next paragraph Distributed resistance Rj; as computed using Equation 5-59(b) The self impedance of the counterpoise r + jwX has been computed using program LINPA (see Chapter 3 and Appendix A) for each value of soil resistivity and current frequency. Rigorously, the computations performed by LINPA are valid for very long overhead conductors. However, the results provide a good estimate of the exact values. Additional information is also provided in Chapter 2. An examination of Figure 5.25 and 5.26 leads to several interesting conclusions regarding the performance of counterpoises during steady- state conditions. Among these are: a- There is an effective counterpoise length, £e, above which the ground impedance is essentially constant. This is equivalent to stating that beyond £e, the current flow in the counterpoise is nearly zero. This effective length is approximately 3 to 5 times l/a, i.e., £e — ^/a. b- The effectiveness of a counterpoise increases as soil resistivity increases. When the modulus of the counterpoise ground impedance is compared with the ground impedance for a perfect conductor, it is clear that long counterpoises provide an effective means of reducing ground impedance in high resistivity soils. c- The performance of a counterpoise deteriorates very quickly as the frequency of the current increases. Figure 5.26 clearly shows that the effective length of the counterpoise decreases rapidly as the frequency increases. This illustrates the effect of transient currents of very short durations. Other phenomena such as soil breakdown and wave reflections also will significantly alter the transient behaviour of the counterpoise. d- As mentioned previously, it is preferable to inject the current from the center of the counterpoise rather than from an extremity. Thus, several counterpoises radiating from a center common point are apparently more effective than a single counterpoise having the same total length and energized from one of its extremities. This conclusion is based on the assumption that the distributed ground resistance is constant over the length of the grounding system. This assumption, although not generally valid for a single linear counterpoise, is a good approximation. Unfortunately, the assumption becomes less and less accurate as the number of conductors radiating from a common center point increases. In this case the resistive coupling between the branches is not negligible and, therefore, should be considered in the computations. As a result, the ground impedance of the counterpoise system will increase as the number of branches is increased, while the total length is kept constant. Although not analyzed in this section, it is reasonable to assume that an optimized counterpoise system of a given length should consist of a number of elements arranged such that the resistive coupling is minimal and the system is energized from several points, suitably located on each element. Unfortunately, the analysis of such a counterpoise system requires considerably more computations than are required for an isolated linear counterpoise. Some typical results from computer solutions of the counterpoise problem are discussed briefly in Section 5.5.3. 5-40 EXTREMITY INJECTION CENTER INJECTION FREQUENCY = 60 Hz C = 0.0 and 0.03 jF/Km 10000 ft-m 10 n-m PERFECT CONDUCTOR COUNTERPOISE LENGTH (meters) Figure 5.25 Ground Impedance of a Counterpoise 5.5.2 COUNTERPOISES OF TRANSMISSION LINES. Transmission line counterpoises do not usually extend beyond the narrow right of way corridor of a line. Thus, continuous conterpoises follow the line and are connected at each structure. Crow-foot (radial) counterpoises have short branches which rarely exceed 50m (150 feet) and thus, can be considered as a concentrated grounding system whose performance can be analyzed according to the method presented in Section 5.3. If the length of the counterpoise branches are too great to permit this approach, the method presented in this section can be used without any modification. 5-^1 p = 100 ft-m Cl = EXTREMITY C = 0.0 and 0.03 uF/Km 500 KHz 100 KHz 50 KHz 10 KHz 5 KHz 1000 Hz 500 Hz 100 Hz .60 Hz 50 Hz 20 Hz PERFECT CONDUCTOR 0.01 Hz COUNTERPOISE LENGTH (meters) Figure 5.26 Effects of Frequency on Counterpoise Impedance In Section 5.5.1 the performance of an isolated continuous counterpoise was investigated. Unfortunatly, the equations which were developed can not be applied to transmission line counterpoises without significant loss of accuracy for the following reasons: a- The per unit length ground resistance of a transmission line counterpoise is not constant over the length of the counterpoise. This results not only from the variations in the mutual ground resistance between the various elements of the counterpoise, but because there is also an additional ground resistance due to the structure foundations or other supplemental grounding elements at each transmission line structure. This situation is illustrated in Figure 5.27. b- When the transmission line is equipped with overhead ground wires, as in most high voltage lines, the counterpoise is energized from multiple points along its length, i.e., at each structure on both sides of the faulted structure (see Figure 5.27). c- The mutual impedance between the counterpoise and the overhead ground wires and the mutual impedance between the phase wires and the counterpoise can not be neglected as they both have a noticeable influence on the performance of the transmission line, including the counterpoise conductor. d- In many cases, there are two counterpoise conductors, one buried on each side of the transmission line. The mutual ground resistance between the counterpoise conductors must be considered, otherwise a significant error will be introduced. 5-42 Phase Conductor Ground Wire ' Counterpoise (a) ACTUAL PROBLEM Transmission Line Structure Groundina (b) Figure 5.27 , Distributed { Resistai Counterpoise EQUIVALENT UNDERGROUND NETWORK Transmission Line Counterpoise Clearly, the concise equations of Section 5.5.1 can not be extended to cover the case of transmission line counterpoises. This is primarily because the distributed resistance R cannot be assumed to be a constant in the wave Equation 5-51 and because there are several injection points along the counterpoise length. Fortunately, an accurate solution can be obtained with the following method which requires that the computations be performed using suitable computer programs such as LINPA, GTOWER, and PATHS, described in the Appendices. These programs are based on the analytical methods presented in this report. This method will now be explained. Since it is possible to accurately analyze the performance of concentrated grounding systems using the equations of Section 5.3, extended grounding systems can also be accurately studied if they are broken down into a number of suitable concentrated grounding system arrangements. In the case of the counterpoise shown in Figure 5.27, this is achieved through the following steps: Step 1 First, it is desirable to separate the effects of the inductive mutual coupling between overhead and buried conductors, from the purely earth conduction effects. Based on the steady-state superposition principle, the counterpoise can be considered to be made of two fictitious conductors: an insulated conductor characterized by self and mutual impedances identical to the original counterpoise values; and a bare perfect conductor, i.e., having zero self and mutual impedances. Thus, the insulated fictitious conductor can be considered as an additional overhead ground wire. Since the second, bare conductor can not experience a voltage drop along its length, it is necessary to make a small opening at the middle of the conductor section connecting the two structures as shown in Figure 5.28. Each section between two consecutive openings will be raised to the potential of the structure to which it is connected. Since this structure is also connected to the middle point of the corresponding section of the insulated, fictitious, counterpoise conductor, this potential is the average potential experienced by the real counterpoise between the opening locations. It is possible to improve the accuracy of this simulation by further subdividing both fictitious counterpoise conductors into smaller elements, regularly interconnected by perfectly conductive jumpers. However, the improvement in accuracy is marginal and cannot justify the significant increase in required computations. 5-h3 Ground Wire Phase Conductor Fictitlous 0/H Counterpoise FICTITIOUS PERFECTLY CONDUCTING COUNTERPOISE Figure 5.20 Equivalent Conductor Pair For a Counterpoise Step 2 Each section of the fictitious buried conductor can now be considered as a concentrated grounding system. The performance of one of these sections can be analyzed using the equations of Section 5.3, provided that sections are separated by distances large enough to justify neglecting their mutual ground resistances. Since this is not the case for this problem, it is imperative to take this proximity effect into account. Two approaches can be used to meet this objective: • The iterative approach • The average proximity effect approach The iterative approach can be as accurate as necessary but requires considerable computation time. In practice, however, such high accuracy is of academic interest only, since the earth structure is very seldom uniform along the length of a counterpoise and several data values are not precisely known. This approach is qualitatively described in step 3a. The average proximity coupling effect is an approximate method which has the advantage of being relatively easy to apply to practical problems. Step 3a 1- Firstly, the mutual ground resistances between the buried sections are ignored and the ground resistance of the sections (TL structure ground resistance) is calculated assuming that each section is isolated from the other. Assume that Rf is the ground resistance of section i. 2- The fault current distribution among the various structure ground resistances R° is determined using the equations developed in Chapter 6, (see Figure 5.29). Computer program PATHS, described in Appendix D, can be used to perform these calculations. 3- A better estimate of the previously determined ground resistance Rf can be obtained by considering the influence of the neighbouring structures j (j = 1, i-1 and j = i + 1, n) which inject a current Ij° into the earth. In practice, only a limited number of structures on each side of structure i need to be considered, since the effect of distant structures can be neglected. A new set of ground resistance values Rf is therefore determined (iteration 1). New structure ground currents If can now be calculated using program PATHS. 4- The iterative process described previously is continued in sequence until the calculated ground resistance R* (or current if ) at iteration k is not significantly different from the computed value at iteration k-1. 5- At this stage (iteration k), the performance of the transmission line counterpoise can be accurately determined from the computation results illustrated in Figure 5-30. For example: The counterpoise ground impedance is: I k f Z = where f designates the structure where the fault occurs and Ig^, Ig2 are the ground wire currents on each side of the fault location. The counterpoise ground potential at each structure i is: The counterpoise ground potential at any intermediate location is determined from the interpolated portion of curve between the computed structure potential points (Figure 5.30(b)) The average longitudinal current in the section of counterpoise between structures i and i+1 is given by the current in the fictitious insulated section between these structures. The average ground leakage current in the section of counterpoise between structures, i , and i + 1 , is: and the ground leakage current at any intermediate point on the counterpoise between the two structures is estimated from the curve joining the points representing one half the ground current at each structure (Figure 5.30(b)). Phase Conductor Ground Fictitious Wire + Overhead Counterpoise Figure 5.29 First Iteration (b) Figure 5.30 CURRENT AND VOLTAGE DISTRIBUTION Last Iteration Step 3b The approach presented in step 3a requires a cumbersome iterative procedure which is only feasible when the soil structure along the counterpoise is reasonably well known and when the transmission line is short; i.e., when the distance between the transmission line terminal substations is relatively short. In most practical cases, however, the structure proximity effect can be estimated using the following procedure: 1- The ground resistance of the total counterpoise length L is calculated using Equation 5-59(a) which assumes that there is no voltage drop along the counterpoise length. Assume that R|_ is the computed resistance. 2- The ground resistance of the portion of counterpoise at structure i (length £) is calculated from 5-59(a). Assume that R.£ is the computed resistance. 3- If there were no mutual ground resistances between the n portions (assumed of equal length i = L/n) of the fictitious counterpoise conductor, we would have: R l nR L i.e., using Equation 5-59(a): np 21 P ^2ae 7T& ttL 21 1 n - ■ ....' -y2ae 1 But this cannot be since L = n £ . Therefore, the actual resistance R £, at each transmission line structure is determined from the relation: i where 7 is a factor (7 > 1) which takes the proximity effect into account. From the two preceding relations, the factor 7 can be written: nR, Y = 2L In — -1 yj 2ae 2£ In -==, -1 yj2ae 5-46 4- Using the corrected resistance values R 7 = Rjj,; the analysis can now proceed as if these values are the values obtained at the final iteration k using the iterative approach described previously. Thus, the final computations are identical to those presented in step 3a, items 4 and 5. 5.5.3 ACCURATE ANALYSIS OF COUNTERPOISES Two of the simplifying assumptions used to derive the counterpoise Equations 5-52 to 5- 56 lead to unrealistic results as frequency and counterpoise length increase. Fortunately, a simple adjustment can be made which will compensate for the inaccuracies introduced by the assumptions. The first assumption is that the value of distributed ground resistance Rj per unit length of counterpoise is uniform along the counterpoise, i.e., independent of the location of the section of counterpoise used to apply the distributed parameter analysis. The second assumption is implicitly made when Equation 5-59(b) is used to calculate Rj. This equation is based on the assumption that current flowing into earth from a cylindrical horizontal ground conductor of length £ is uniformly distributed along the conductor. As previously discussed in this chapter and in [14,15,21], the current density is generally significantly higher at the conductor extremities than at the center. These two simplifying assumptions are interrelated because of the necessity to accurately calculate the conductor current distribution and earth potential distribution along the surface of a ground conductor, in order to determine the ground impedance of the conductor as seen from the point where the current is injected into the conductor. This kind of calculation can only be carried out realistically using a computer program such as the SES program MALZ [45]. Figure 5.31 shows the ground impedance of a counterpoise as a function of its length based on the assumption that: • The counterpoise is a perfect conductor (zero internal impedance). In this case, Equation 5-59(a) applies. This was used to generate the solid curve of Figure 5.31. • The counterpoise is made of small identical sections which can be analyzed using the distributed parameter approach (Equations 5-52 to 5-59). • The counterpoise is made of a number of sections having a uniform surface potential and earth current density on each section but not necessarily the same from one section to the other. The potentials and currents of the sections are determined by MALZ with the desired accuracy by subdividing the counterpoise into an appropriate number of short sections. Examination of Figure 5.31 reveals that as the length of the counterpoise is increased, the ground impedance decreases, as expected, until a certain length is reached, after which the impedance rises. This sudden impedance increase has no physical justification but is simply the result of approximations on which the counterpoise resistance Rj is based. The MALZ program generated curve also shows a similar but less pronounced behavior. This effect virtually disappears as the number of counterpoise sections increases. The optimum number and length of counterpoise sections is frequency and resistivity dependent. It is interesting to note that the computer results consistently show that when the number of sections is sufficient, the counterpoise ground impedance curve becomes almost horizontal beyond the minimum impedance point on the distributed parameter curve (knee point). This suggests that the distributed parameter curve can still be used, provided that the rising portion of the ground impedance curve is ignored and replaced by a horizontal line. The 5-47 ground impedance curves of Figure 5.26 and 5.25 and those shown in Volume 2 were obtained after this adjustment technique was applied. Another difficulty associated with the distributed parameter approach arises when the counterpoise current distribution is to be determined. Based on the Equations 5-52 and 5-53, the current is proportional to the hyperbolic sine of the distance from the structure to the location where the current is to be observed. This is different from the current density curves of Figure 5.32 which are based on MALZ computations. This figure clearly shows that the current distribution is dependent on the length of the counterpoise and that, for short counterpoises, an end effect consisting of an increase in current density can be observed. A closer examination of this figure and of Figure 4.17 of Volume 2 also reveals that the current distribution and end effects are frequency and resistivity dependent. An accurate calculation of current distribution along isolated counterpoises is of limited interest in transmission line grounding problems. The so-called continuous counterpoise most common in transmission line applications is buried in various soils and is regularly connected to the structures and the overhead ground wires. Consequently, it is more effectively and accurately analyzed using the methods described in Section 5.5.2. P CI - 100 {2-m (CURRENT INJECTION) = EXTREMITY DISTRIBUTED PARAMETERS -SES COMPUTER PROGRAM MALZ - assuming zero IMPEDANCE COUNTERPOISE __ 60 Hz COUNTERPOISE LENGTH (meters) Figure 5.31 Computer Solutions of Counterpoise Ground Impedance 5-48 LINEAR CURRENT DENSITY ALONG COUNTERPOISE (A /m ) £ = 10000 m p = 100 ft-m FREQUENCY = 60 Hz CI = EXTREMITY 8000 m 6000 m 2000 m 1000 m 800 m ,—600 m 200 m 100 m 20 m t—i" i i [ COUNTERPOISE LENGTH (meters) Figure 5.32 Current Distribution Along a Counterpoise 5.6 SOIL HEATING EFFECTS The importance of soil moisture content has been discussed on several occasions in this chapter and previously. It is now clear that a ground electrode should not be loaded continuously, or even for short periods, without suitable provision for heat dissipation in the earth. This is necessary for maintaining the temperature rise of the soil in the vicinity of the electrode, safely below the 100 °C limit of water evaporation. Fortunately, it is very unlikely that the duration or magnitude of high voltage transmission line fault currents will exceed the maximum value above which soil mixture evaporation becomes a real concern. Hence, the subject of soil heating will not be analysed in detail. Nevertheless, there does exist the possibility that a fault occurs at a high resistance transmission line structure not equipped with overhead ground wires or other zero-sequence return paths. Generally, low voltage transmission or distribution lines are the most prone to this danger, especially, when the support structure is a concrete pole with no supplemental grounding elements [40,41]. In this situation, the fault current may be low enough to prevent the operation of the protection system or, a long time may elapse before a protective element is activated. Consequently, a sustained current may penetrate the soil via a relatively small ground electrode surface. The resulting high current density will introduce localized heating and a corresponding increase in soil resistivity. This, in turn, will cause a further increase in the local soil temperature. This cumulative effect, may lead to a complete evaporation of the moisture in the soil around the grounding system, followed by destructive arcing and, ultimately, complete isolation of the grounding system from remote ground [40]. This process may take a few minutes to several hours or days, depending on the fault current magnitude, earth structure, grounding system configuration and ambient temperature. The following theoretical considerations will provide a means of assessing the seriousness of the heating process, should it arise during the design stage of a transmission line project. It will allow for an evaluation of alternatives in the design of an existing structure grounding system subject to such heating stresses. 5.6.1 HEATING OF THE SOIL AROUND A GROUNDING SYSTEM Since the steady-state heat and current flow obey the same mathematical law, namely, Poisson's differential equation, the heat performance of any grounding electrode can be determined from a knowledge of the earth electric field , provided that certain conservative assumptions are made: ® The atmosphere is a perfect insulator. • The electrode surface is an isothermal surface, i.e., there is no temperature drop between any two points on the electrode surface. From the preceding discussion, it follows that the temperature rise around a grounding electrode can be determined by considering the temperature rise of a hemispherical electrode where terms related to the electric field are replaced with values corresponding to the grounding system under consideration [35,46-48]. In this section however, the analysis is conducted for an arbitrary grounding system. Figure 5.33 shows a ground electrode injecting a current I into a uniform soil. The differential equation of the temperature rise in the soil is [35,47]: 36 Y------ X&0 = Q (5-60) 3t 5-50 where 6 is the temperature rise at any point in the soil X is the thermal conductivity of the soil (assumed constant) y is the specific heat of the soil (assumed constant) Q is the energy (in joules) produced per second and per cubic meter A is the Laplacian operator The following average values of y y and X are provided in [46]: ~ 1.5x106 J/m^.°C X ~ 2.0 /m.°C EQUI POTENTIAL AND ISOTHERMAL SURFACES Figure 5.33 Heating of the Soil Around a Grounding Electrode The solution of Equation 5-60 is cumbersome, even for simple electrode configurations such as the ground rod. Fortunately, a complete solution is necessary only when an accurate knowledge of temperature variation with time is critical. In problems involving the heating of soil by grounding electrodes, the results of practical interest are the steady-state temperatures and sometimes, the initial (transient) temperature variations. Note that the transition period between the transient and steady-state temperatures can be approximated by interpolation, (see Figure 5.34). 5.6.2 STEADY-STATE PERFORMANCE During the stationary condition, the time derivative term of Equation 5-60 vanishes and this equation, expressed in cartesian coordinates, reduces to: 320 020 320 E2 9x2 3y2 9z2 Xp (5-61) 5-51 where E is the electric field intensity (Q = E2/p) P is the average earth resistivity Recalling that the electric field intensity components are: 3U su — j 3x E = - ----y By * 3U E = - ----- z 3z where U is the earth potential at an arbitrary point, and since AU = 0 is the equation of the earth potentials, the solution of Equation 5-61 is given by: 9 = (AU - iU2)Ap (5-62) where A is a constant. Based on these assumptions, there is no temperature drop between any two points on the electrode surface. Thus, on the surface, U = Us and therefore: 90 90 90 — = — = — = o 9x 9z By a condition which, after differentiation of 5-62, leads to A = Us. The following remarkably simple equation is thereby obtained for the temperature rise of the soil at the surface of a grounding system, regardless of its shape: U2 R2!2 (5-63) 0 = — = ------2Ap 2Ap where R is the ground resistance of the electrode. Thus, the maximum continuous current which can flow in the grounding system without exceeding the allowable temperature rise ^ is: 1 (5-64) = - j2Ap6m Assuming a maximum temperature of 100 °C, it is straightforward to conclude that the allowable continuous current density of an electrode is relatively small, i.e., in the order of few amperes per m2? considerably less current than the 5 A/cm^ considered in Section 5.4 for short duration currents. 5.6.3 TRANSIENT HEATING During the initial period of current conduction into the soil, the heat does not have the time to dissipate effectively into the soil by conduction. Thus, the thermal conduction term in Equation 5-60 can be neglected with regard to the specific heat term. This approximation is, of course, a conservative one. Equation 5-60 reduces to: de o. e2 dt y YP 5-52 therefore (5-65) The maximum temperature rise occurs at the point of maximum field intensity Em, usually a very short distance from the electrode surface. The electric field intensity E and the earth current density J at an arbitrary point are related according to the expression: E = pJ Consequently Equation 5-65 can be written as: Vy 0 (5-66) 7 7 This equation can be used to determine the maximum short duration current density which does not result in a temperature rise exceeding a safe value. For a time duration of 0.1 s, soil resistivity of 100 ^-m, and a temperature rise of 80 °C, the maximum density is: I.SxlO6 80 ------------x ------ V 100 - 31(00 A/m2 - 0.3** A/cm2 0.1 This value is significantly lower than the 5 A/cm2 short duration current limit in concrete foundation specified in Section 5.4. Since the value 5 A/cm2 is based on experimental evidence [44], it is believed to be realistic. This is true provided that sufficient time passes between conduction periods at the electrode, to permit moisture migration from neighbouring soil back into the areas affected by water evaporation. Figure 5.34 illustrates the variation of temperature with time for short duration and steady-state currents (solid line). For intermediate times, the interpolated portion of the curve provides a good estimate of temperature variations. Iransient Figure 5.34 Steady State Temperature Variation With Time 5-53 The time constant of heating r, i.e., the time required to reach the steady-state temperature, is obtained (approximately) by equating 5-62 with 5-65. Thus: u2 (5-67) T 2E2 The time constant at the surface of a grounding electrode (U=US E=ES) is: _ 1 Y Ts = I 7 For a hemispherical electrode of radius U pi = -----5 2na ; and E a: pi = ------5 2ira2 hence Y T S a2 2\ This indicates that the time constant of a grounding electrode is a function of the square of its average dimension. For a 5 m radius electrode, the time constant is: 1.5x106 = ----------- x 25 - 107 seconds ~ r 5 108 days 1 k Although this is a long period, experience shows that periods in the order of 10 to 100 days are typical values for practical grounding systems. REFERENCES 1 - T. N. Giao. M. P. Sarma, "Effects of two-Layer Earth on the Electric Field Near HVDC Electrodes", IEEE Transactions, PAS-91 No. 6, November 1972, pp. 2346 to 2355. 2 - F. Dawalibi, D. Mukhedkar, "Ground Electrode Resistance Measurements Nonuniform Soils", IEEE Transactions, Vol. PAS-93, No. 11, January 74, pp. 109-115. in 3 - A. I. Yacobs, P. I. Petrov, "On Allowing for the Longitudinal Impedance of Horizontal Elements in Large Earthings", Electrical Technology USSR (GB) No. 1, 1974, pp. 9-22. 4 - F. Dawalibi, D. Mukhedkar, "Optimum Design of Substation Grounding in Two-Layer Earth Structure - Part I, Analytical Study", IEEE Transactions, Vol. PAS-94, No. 2 March/April 1975, pp. 252-261. 5 - F. Dawalibi, D. Mukhedkar, "Optimum Design of Substation Grounding in Two-Layer Earth Structure - Part II, Comparison Between Theoretical and Experimental Results", IEEE Transactions, Vol. PAS-94, No. 2, March/April 1975, pp. 262-266. 5-54 6 - F. Dawalibi, D. Mukhedkar, "Optimum Design of Substation Grounding in Two-Layer Earth Structure - Part III, Study of Grounding Grids Performance and New Electrode Configurations", IEEE Transactions, Vol. PAS-94, No. 2, March/April 1975, pp. 267-272. 7 - J. G. Sverak, "Optimized Grounding Grid Design Using Variable Spacing Technique", IEEE Transactions, Vol. PAS-95, No. 1, January/February 1976, pp. 362-374. 8 - A. M. Arakelian, R. O, Gabrielyan, D. V. Arytyunyan, "Quick Calculation of Parameters of Earthing Rods in Inhomogeneous Media", Electrical Technology USSR (GB), No. 3, 1976, pp. 58-61. 9 - F. Dawalibi, D. Mukhedkar, "Multi-Step Analysis of Interconnected Grounding Electrodes", IEEE Transactions, Vol. PAS-95, No. 1, January/February 1976, pp. 113-119. 10- F. Dawalibi, D. Mukhedkar, "Resistance Calculation of Interconnected Grounding Electrodes", IEEE Transactions, Vol. PAS-96, No. 1, January/ February 1978, pp. 59-65. 11- F. Dawalibi, D. Mukhedkar, "Transferred Earth Potentials in Power Systems", IEEE Transactions, Vol. PAS-97, No. 1, January/February 1978, pp. 90-101. 12- R. J. Heppe, "Step Potential and Body Currents for Near Grounds in Two-Layer Earth", IEEE Transactions, Vol. PAS-98, No. 1, January/February 1979, pp. 45-59. 13- H. N. Nunnally, R. P. Webb, E. B. Joy and A. P. Meliopoulos, "Computer Simulation for Determining Step and Touch Potentials Resulting from Faults on Open Neutrals, in URD Cable", IEEE Transactions, Vol. PAS-98, No. 3, May/June 1979, pp. 1130-1136. 14- F. Dawalibi, D. Mukhedkar, "Influence of Ground Rods on Grounding Grids", IEEE Transactions, Vol. PAS-98, No. 6, November/December 1979, pp. 2089-2098. 15- F. Dawalibi, D. Mukhedkar, "Parametric Analysis of Grounding Grids", IEEE Transactions, Vol. PAS-98, No. 5, September/October 1979, pp. 1659-1668. 16- F. Dawalibi, D. Mukhedkar, "Resistance Measurement of Large Grounding Systems", IEEE Transactions, Vol. PAS-98, No. 6, November/December 1980, pp. 2348-2354. 17- D. L. Garrett and H. J. Holley, "Calculation of Substation Grounding Resistance Using Matrix Techniques", IEEE Transactions, Vol. PAS-99, No. 5, September/October 1980, pp. 2008-2011. 18- F. Dawalibi, W. G. Finney, "Transmission Line Tower Grounding Performance in Nonuniform Soil", IEEE Transactions, Vol. PAS-99, March/April 1980, pp. 471-479. 19- P. Kouteynikoff, "Numerical Computation of the Grounding Resistance of Substations and Towers", IEEE Transactions, Vol. PAS-99, No. 3, May/June 1980, pp. 957-965. 20- A. P. Meliopoulos, R. P. Webb, E. B. Joy, "Analysis of Grounding Systems", IEEE Transactions, Vol. PAS-100, No. 3, March 1981, pp. 1039-1048. 21- F. Dawalibi, D. Mukhedkar, D. Bensted, "Measured and Computed Current Densities in Buried Ground Conductors", IEEE Transactions, Vol. PAS-100, No. 8, August 1981, pp. 4083-4092. 22- T. Takashima, T. Nakae, R. Ishibashi, "High Frequency Characteristics of Impedances to Ground and Field Distributions of Ground Electrodes", IEEE Transactions, Vol. PAS-100, No. 4, April 1981, pp. 1893-1900. 5-55 23- D. Bensted, F. Dawalibi, A. Wu, "The Application of Computer Aided Grounding Design Techniques to a Pulp and Paper Mill Grounding System", IEEE Transactions, Vol. IA-17, No. 1, January/F ebruary 1981. 24- D. S. Ironside, "Some Recent Developments in Portable Ground Resistance Test Instruments", Paper presented at the High-Voltage Power System Grounding Workshop Sponsored by EPRI, May 12-14, 1982, Atlanta, Georgia. 23- I. D. Lu, R. M. Shier, "Application of a Digital Signal Analyzer to the Measurement of Power System Ground Impedances", IEEE Transactions on PAS, Vol. PAS-100, April 1981, pp. 1918-1922. 26- P. H. Reynolds, D. S. Ironside, A. H. Silcocks, J. B. Williams, "A new Instrument for Measuring Ground Impedances", IEEE Paper A79 080-3, IEEE PES Winter Meeting, New York, 1979. 27- J. Ufermann, K. Jahn, "High Frequency Earth Resistance Measuring Instrument", Brown and Boveri Journal No. 3/67, pp. 132-135. 28- M. Darveniza et Al, "Modelling for Lightning Performance Calculations", IEEE Transactions on PAS, Vol. PAS-98, No. 6, November/December 1979, pp. 1900-1907. 29- AIEE Committee Report, "A Method of Estimating Lightning Performance of Transmission Lines", AIEE Transactions on PAS, Vol. 69, pt.II, 1950, pp. 1187-1196. 30- Report of Joint IEEE-EEI Subject Committee on EHV Line Outages, IEEE Transactions on PAS, Vol. PAS-86, 1967, p. 547. 31- E. Clarke, "Circuit Analysis of AC Power Systems, Volume 1; Symmetrical and Related Components", Book Published by General Electric Company, 1961. 32- "Transmission Line Reference Book 345 kV and Above", General Electric Co., Published by EPRI. 33- M. Muskat, "Potential Distribution About an Electrode on the Surface of the Earth", Physics, Vol. 4, No. 4, April 1933, pp. 129-147. 34- H. M. Mooney, E. Orellana, H. Pickett, L. Tornheim, "A Resistivity Computation Method for Layered Earth Models", Geophysics, Vol. XXXI, No. 1, February 1966, pp. 192-203. 35- E. D. Sunde, "Earth Conduction Effects in Transmission Systems", Dover Publications, New York, 1968 (book). 36- E. A. Cherney, K. G. Ringler, N. Kolcio, G. K. Bell, "Step and Touch Potentials at Faulted Transmission Towers", IEEE Transactions on PAS, July 1981, Vol. PAS-100, No. 7, pp. 3312-3321. 37- G. B. Niles, "Using Transmission Foundation Resistance for Grounding Purposes and Determining an Effective Earth Resistivity", IEEE Paper A78 126-5, IEEE PES Winter Meeting, New York, February 1978. 38- M. Datta, A. K. Basu, M. M. R. Chowdhury, "Determination of Earth Resistance of Multiple Driven-Rod Electrodes", Proc. of the I.E.E, Vol. 114, No. 7, July 1967, pp. 1001-1007. 39- S. A. Sokolov, "The Lowering of Earth Resistances in Rocky Areas and Permafrost Regions" Electric Technology USSR (GB), Translation of Electrosvyaz, No. 6, 1959, pp. 65-70. 5-56 40- W. Bogajewski, F. Dawalibi, Y. Gervais, D. Mukhedkar, "Effects of Sustained Ground Fault Current on Concrete Poles", IEEE Paper 82 WM 202-1, IEEE Winter Meeting, Atlanta, 1981. 41- W. K. Dick, H. R. Holliday, "Impulse and Alternating Current Tests on Grounding in Soil Environment", IEEE Transactions, PAS-97, No. 1, January/F ebruary 1978, pp. 102-108. 42- E. J. Fagan, R. H. Lee, "The Use of Concrete-Enclosed Reinforcing Rods as Grounding Electrodes", IEEE Transactions, Vol. Iga-6, No. 4, July/August 1970, pp. 337-348. 43- J. Preminger, "Evaluation of Concrete-Encased Electrodes, IEEE Transactions, Vol. IA-II, No. 6, November/December 1976, pp. 664-668. 44- R. Kostaluk, "Effects in Concrete and Ground Caused by Passage of Lightning and Short-Circuit Currents Through Foundations of Transmission Lines", Archiwum, Elektrotechniki, Tom XXV, 1976, pp. 95-107. 45- F. Dawalibi, "User Manual for Program MALZ", Safe Engineering Services Ltd., Montreal, 1981. 46- E. W. Kimbark, "Direct Current Transmission" Volume I, Wiley Interscience, 1971 (book). 47- J. H. Neher, "The Temperature Rise of Buried Cables and Pipes", AIEE Transactions, Vol. 68, Part I, 1949, pp. 9-21. 48- K. R. Rakhimov, "Thermal Analysis of Earthing Rods at the Ends of D.C. Lines". Electric Technology USSR (GB), Vol. 4, July 1970, pp. 36-43. 5-57 FAULT CURRENT DISTRIBUTION 6.1 GENERAL The majority of transmission line faults are to ground and generally occur between a phase conductor and a transmission line structure, often, as the result of an insulator flashover. In some cases, the presence of foreign objects between a phase conductor and the overhead ground wire (birds, etc.) or a grounded structure (trees or the arm of a crane) may cause a ground fault somewhere along one span of the transmission line. Occasionally, the ground fault is caused by a broken phase conductor in direct contact with the overhead ground wire or soil surface. In all these cases, the fault current return paths include earth and, therefore, present an impedance which at most is equal to the equivalent ground impedance of the grounded structures which carry the fault current (the ground impedances of the generating sources are neglected). If in addition to the earth current, part of the total fault current returns to the generating sources via metallic return circuits (such as overhead ground wires, counterpoises, etc.) the impedance value will be even less. In conventional short circuit calculation techniques, the ground impedance is usually neglected by assuming a zero impedance return path. Because modern power systems are solidly grounded and the current return path impedance is generally negligible with respect to the other power network impedances, this zero impedance return path assumption is very accurate for most engineering problems and leads to conservative results. When this assumption is not realistic, it is possible to add the value of the return path impedance in the equivalent power system circuit. This procedure allows the influence of the ground impedance on the magnitude of the total fault current to be included in the calculations (see Figure 6.1). In all these cases, the division of the fault current at the faulted structure between earth and the metallic return conductors of the circuit remains unknown. This is because of the simplification of the circuit resulting from the use of an equivalent return path impedance (generally, an approximated conservative impedance value). In problems where a proper knowledge of the distribution of the fault current among various paths is necessary, the conventional short-circuit computation tools are inadequate and alternative methods must be used. This chapter describes several of these alternative methods. One of these methods was used as the basis for the development of the computer program PATHS described in Appendix D. The subject of transmission line ground fault current distribution is not extensively covered in the technical literature. However, the existing work related to this subject [1-10] is comprehensive, reasonably complete and reflects the two principal approaches used to analyze the problem, i.e.: • Transmission Line with constant line parameters • Transmission Line with varying line parameters Z z r Z s se Z s + Zr Conventional Equivalent Circuit Actual Circuit Figure 6.1 Equivalent Return Path Impedance Conventional Circuit Reduction Selection of the most suitable computation technique depends on the type of problem being analysed, the availability and accuracy of the input data and the required accuracy of the results. Sometimes, simple hand calculations are all that is required to analyze the problem. In other cases, complex computer programs are required to arrive at meaningful conclusions. In this report, simplified and more complex calculation techniques are briefly discussed and properly referenced. Particular emphasis is placed on the description and discussion of the double-sided elimination method which is suitable for solving the problem of transmission line fault current distribution in general. This method was used as the basis of the computer program PATHS to address the widely variable needs of North American Utilities (see Appendix D). 6.2 IMPORTANCE OF FAULT CURRENT DISTRIBUTION In assessing the effects of a ground fault on a transmission line, it is apparent that the transmission line performance at the faulted structure and/or at other locations on the transmission line, is significantly influenced by the fault current distribution between the structure and the neutral conductors connected to the structure. Figure 6.2 shows a typical system which illustrates this aspect. R represents the faulted structure ground resistance and Zs the equivalent power generation (source) impedance. It is also assumed that the value of Zs is significantly higher than the structure resistance value R so that the total fault current Ij ~ V/Zs is approximately constant when R varies within a relatively narrow range. Inspection of Figure 6.2 leads to the following conclusions: a- If one of the ground wire impedances Z^ or Zr is smaller relative to R, then very little current enters the faulted structure and the potential rise of the structure is proportional to the metallic return circuit path impedance, i.e. Z^ in parallel with Zr. b- If in contrast, R is small compared to Z^ and Zr, then most of the fault current flow through the structure, causing a ground potential rise proportional to the structure ground resistance R. 6-2 This example, although simple and perhaps unrealistic, illustrates the importance of fault current distribution. In a real transmission line there are hundreds of grounded transmission structures. The metallic return circuits are not only bonded to these structures but also are electromagnetically coupled to the phase conductors. Moreover, the ground resistance of the structures is not constant and generally varies between wide limits along the route of the transmission line. The fault current distribution is therefore more complex to calculate and will vary with the type of transmission line and location of the fault on the line. Z s Overhead Ground-wire Fau1 ted Structu re Figure 6.2 Fault Current Distribution 6.3 CONSTANT LINE PARAMETERS One of the most classical methods using uniformly distributed parameters was presented by Endreyni in 1967 [1]. Alternative approaches have since been proposed [3-5], but in essence, the techniques proposed are similar to Endreyni's method or to Rudenberg's earlier but more approximate method [2]. In all cases however, the final equations developed by the authors are valid only for "long" transmission lines. 6.3.1 LONG TRANSMISSION LINES The full treatment of this concept is well covered by Desieno et al [3]. In this section, the case of a ground fault fed from both terminals of a transmission line is described. Figure 6.3 represents the transmission line with the following uniformly distributed parameters: Zg: Self impedance of the ground wire (or equivalent single ground wire); in ohms per unit length (km or mile). Zm: Average mutual impedance between a phase conductor and the (equivalent) ground wire in ohms per unit length. R : Average structure footing ground resistance in ohms. N : Number of transmission line structures per unit length. 6-3 11 If 0 i(x+dx) Figure 6.3 Distributed Parameters Method It is assumed that the fault occurs at some distance from the terminals so that the transmission line can be considered infinite in both directions. This is required to maintain the validity of the distributed parameters approach. If this is not the case, then the equations of this section should not be used. Instead, one should use the equations of the method developed in the next section (6.3.2), or those related to methods based on varying line parameters (see Section 6.4). The length of the three-phase single ground wire transmission line of Figure 6.3 is v. A phase to ground fault occurs at a distance d from the left terminal. The current i in a transmission structure located at a distance x from the left terminal is: -j---- = i (x) = l(x + 6) - l(x) (6-1) where I(x) is the ground wire current in the local span and I(x+ 5) the ground wire current in the next span (see Figure 6.4). 5 represents the distance between consecutive transmission line structures. Because a very long transmission line is assumed, the average ground resistance R (ohms) of the N structures per km or mile of the line can be reliably approximated by a large number 7/ of shunt resistances of R' where, R' = y R/N. The distance between any two consecutive resistances is dx = I/17. Rewriting Equation 6-1 for this new but equivalent line configuration leads to: V(x) R1 Ndx V(x) = I(x + dx) - I (x) Dividing both sides of the equation by dx and taking the limit of the right-hand side as dx approaches zero leads to: H V(x) = ■d.I(x). R W dx (6-2) The total drop of voltage in the circuit consisting of two shunt resistances R' and the short segment of ground wire (of length dx) between these resistances must be zero (see Figure 6.4): V (x \ + dx) - V(x) - l(x)Z g dx + f m dx = 0 Irl 6-4 Dividing by dx and taking the limit of the preceding eguation as dx approaches zero yields the following relation: Z I(x) - Z l' g m f = dx (6-3) By differentiating (6-2) and (6-3) with respect to x and then combining the resulting equations together, the following second order differential equations are obtained: d2l(x) - cc2l(x) = - a2 ^ If (6-4) g dx" d2V(x) - a2V(x) = 0 (6-5) NZ ■t (6-6) dx" where a = V The general solutions of (6-4) and (6-5) are respectively: (x) = ax -ax , m A„e~"' + B„e l* + Z ‘f (6-7) g „ z v R A ax VX) = N “V R _ -ax (6-8) where the subscript i is used to designate the left terminal. The A and B constants are determined from the boundary conditions, i.e.: At the Faulted structure (x = d) (6-9) Vx * d> - RIe At the Left Terminal (x = 0) It is assumed that the ground resistance of the left terminal is negligibly small. Therefore: (6-10) V (x = 0) = 0 i The solution of the boundary conditions 6-9 and 6-10 yields: NI A£ ~ B£ ~ 2a sinh(ad) Thus sinh(ax) Vn = RI £ e (6-11) sinh(ad) NI Vx) * — a cosh(ax) sinh(ad) * m T ^!f (6-12) g 6-5 The last three equations are related to the left terminal. Similar expressions can be obtained for the right terminal (replace x by y, d by v - d and If by If ), i.e.s A NI = --------------------------- = B r r 2a sinh[a(v-d)] sinh(ay) V (y) = RI -------------------r e sinh[a(v-d)] NI ir(y) ■ — cosh(ay) - (6-13) Z . ^ r ‘f (6-14) At the faulted structure (x = d ; y = v-d), the total line to ground fault current If is: If = If + If = Ij, + Ir + Ie From this equation and Equations 6-12 and 6-14, the structure fault current is determined: (1-Zm/Zg)lf I e = 1 ^ (6-15) 1 tanh(ad) tanh[a(v-d)] The current I in the ground wire and the structure potential rise V (or structure current i = V/R) at any point along the line are determined from Equations 6-11 to 6-14 by substituting Ie by its value as given from Equation 6-15. Equation 6-15 shows that the current Ie which flows in the faulted structure is a fraction of the total fault current If. This current decreases when: • The mutual impedance between the phase conductor and the ground wire increases. • The ground wire impedance decreases. • The structure resistance increases. From Equation 6-15, it is clear that the value of this structure current Ie can not exceed (1 - //) times the total fault current. ^ is the coupling factor and is equal to the ratio of the mutual impedance Zm to the ground wire impedance Zg. Depending on the material used and number of ground wires installed, the value of pt generally varies between 0.1 and 0.5. It is also apparent from Equations 6-12 and 6-14 that, in the presence of coupling between the phase and metallic return conductors, a portion of the current (julf or jttlf) will stay in the metallic return circuits ("trapped" current). 6.3.2 LONG AND SHORT TRANSMISSION LINES The preceding method can not be applied for short lines, i.e., lines consisting of few spans between the feeding station and the faulted structure. In this case the method of lumped parameters should be used. 6-6 When it is assumed that the transmission line conductor characteristics and the structure ground resistance are constant, then the lumped parameter method becomes very similar to the distributed parameter method. That is, instead of the differential equations, difference equations are used to derive the fundamental relations. To investigate this case, the transmission line of Figure 6.4 is considered. The fault occurs at structure number 0. There are n - 1 structures between this structure and the left terminal and m - 1 structures between structure 0 and the right terminal. All structures have a ground resistance R except for structure 0 whose resistance is Re. The left and right terminal ground resistance values are R^ and Rr respectively. The impedance of the section of ground wire between the terminal and the first structure (structures n-1 or m-1) is Z£ for the left terminal and zr for the right terminal. The self and mutual impedances of a section of ground wire between two consecutive structures are Zg and Zm respectively (expressed in ohms). i if f n-1 . . . j +2 j+1 . .2 Figure 6.4 Lumped Parameters Method The current ij+^ in a transmission structure j+1 is: (6-16) 'j+1 " Vi “ Ij+2 where Ij+^ is the current in the section of ground wire between structures j and j+1. According to Ohm's law, Equation 6-16 can be rewritten as: V. V. J+1 J +2 R where Vj is the potential rise of structure j and Zg is the ground wire impedance between two consecutive structures. Rearranging this equation yields: V. J +2 <2 * h Vi * V 0 (6-17) 6-7 According to Kirchoff's law, the total voltage drop in a loop is zero. Therefore: I . ,R + Z I, =0 j+1 m f j+i g which, based on Equation 6-16, can be rewritten as: I.R - I.^Z j V ' <2 ^ = -t1) (6-18) (6-19) Equations 6-17 and 6-19 are second order difference equations which hold for j = 2, 3,------,n-2. That is, they apply at all but nodes 1 and n-1, where the following boundary equations are applicable: At the faulted structure (6-20) -V2 ♦ <2 ♦ -f) V, - V0 - I.R e e At the Left Terminal Z Z 11 (1 + -2- + -!■) v -V . = -2- V = -2- i R„ z^ R n-1 n-2 z^ n z^ n £ (6-21) Finally, at the left terminal, node n, the following relation must be verified: (6-22) I + i = I£ n n f but since I = n V .+ V Z -iTJ---D. + JE z£ Z£ , V , R„ Z , ' = _niL + _A ; + JE x f Zl ZSL n ZZ f Equation 6-22 yields z„-Z i n = 1r f VR£ z£+R£ (6-23) From Equations 6-21 and 6-23, the following equation is obtained after elimination of in; (1 r -i + f )Vn- - V - 1 'n"2 ■■ V Ri R,I,(I - A 2 f n (6-24) The complete solution of the second order homogeneous difference Equation 6-17 is: \ VS,(k) Ag,e ak , „ -ak + B£e (6-25) where a = 2sinh~1( ) (6-26) and the subscript £ is used to designate the left terminal. The A and B constants are determined from the boundary Equations 6-20 and 6-24. The following equations are obtained after some simple algebraic transformations: 6-8 ,D T w - (n-1 )cx,, a, AR£Xf " V0e (6-e ) AA“ (6-27) e(n-l)a(b.e-a) _ e-(n-1)a(b_eaj (6-28) B£ = Vo " A£ where A = ) -£ 0 ~ f-) ; a = ~ + 2 ; b = i 1 + ^ + J— £ £ The ground wire current Ij and the structure potential rise are related according to the following equation: I. = (V - V )/Z + zil/zn J J 1 J 9 m f g Therefore, the complete solution of the nonhomogeneous difference Equation 6-19 is: *f [ Vk> a DA^ ♦ (ea-1 )B^e-ka + ZJ L ;] (6-29) A similar equation is also obtained for the right terminal: Lr{k) [<«-“- l)Arak“ + (eC,-l)Bre’k“ -ka - 2^,] j ve - i;Mre f -1)Or( (6-30) 'g The Ar and Br coefficients are also given by Equations 6-27 and 6-28 in which R^, I £ and n are replaced by Rr, Ir and m respectively and A is defined as: Zn A = (----- 2__) Z d _ _D1) r r r At the faulted structure the total line to ground fault If is: If = If + If = I£(1) + Ir(D + Ie This last equation and Equations 6-29 and 6-30 yield: Z (1 - j1) If + (U£ + Ur)S I e JL = (6-31) 1 + “ (2 " W£ - Wr) where . S _ a = e - e -a ui = x Wj = ra<n-2)“(b-e‘a) - a-(n-2)a(b-e“) /T, = e(n“l)a(b-«'“) - e'(n'1)“(b-e“) 6-9 and Ur? Wr and Tr are defined based on relations similar to those used to define U£, and l£ . It should be noted that while Ie can easily be determined from Equation 6-15 using hand calculations or conventional hand-held calculators, the faulted structure current I@ based on equation 6-31 can only be reasonably calculated using a sophisticated programmable calculator or a computer. If the terminal ground impedances are assumed to be negligible (R^ ~ Rr ~ 0), the preceding equations simplify considerably. In this case a is equal to b and Equations 6-25, 6-29 and 6-31 reduce to the following: s i nh[(n-k)al V (k) = I R ---------------------6 e sinh(na) R Vk) = ^ I Z 9 + (6-32) , 731 9 (6-33) If (6-34) e [<}>(1 ,n) + (p(l ,m) ] where sinh[(n+1-k)a] - sinhC(n-k)a] (6-35) <j>(k,n) sinh(na) Since Equation 6-34 is very similar to Equation 6-15, a comparison between the results of the computations using these equations yields interesting information concerning their capabilities and limitations. Figure 6.5(b) shows in graphical form the computation results based on Equation 6-34 (solid lines) and Equation 6-15 (broken lines). The transmission line which was analyzed is shown in Figure 6.5(a). This figure provides all the data used to perform the calculations. Similar curves are also shown in Volume 2 of this report. Firstly it should be noted that the lumped method results coincide with those obtained using other exact methods such as the generalized double-sided method described later in this chapter. The distributed method yields a faulted structure current value lower than the value predicted by the lumped method. This is true regardless of the fault location. However, the difference between the distributed and lumped method results decreases as the distance from the terminal to the faulted structure increases. At a distance of 3 to 5 km (2-3 mi) from the terminal, this difference remains constant and independent of the fault location (about 10%). The improvement of the accuracy is logical since the distributed method is valid only when the distance between the faulted structure and the terminal is sufficiently long (5 km or more) so that the discrete structure resistances can be replaced by distributed resistances. When the fault is at the second structure from the terminal, the section of line is too short and the computations yield an error of about 33%. 6-10 The 10% minimum error which remains regardless of the fault location, is inherent to the equations applicable to the distributed and lumped methods, as shown overleaf. f Z =2.28+j0.77 tt/km d Terminal LT variable v = 30 km (a) REFERENCE TRANSMISSION LINE Terminal RT Based on the Lumped Parameters Concept Based on the Distributed Parameter Concept Terminal (LT) FAULTED STRUCTURE NUMBER (^2) (b) DISTRIBUTED VERSUS LUMPED CONCEPTS Figure 6.5 Distributed and Lumped Method - Constant Structure Resistance 6-11 When the fault occurs at large distances from both terminals, Equations 6-15 and 6-34 reduce to the following similar equations. These new equations were obtained by replacing d, (v-d), n and m with f and taking the limit of the original expressions as f increases to infinity: - Distributed : I e - Lumped : i e = (l-Z /Z )l,/(l + 2 JrN/Z ) mgr 9 = (l-Z /Z )lf/coth(~) mgr 2 ; a = Zsinh'1 (Q.5\!z /RN ^ 9 ) where Zm and Zg are expressed in ohms/unit length. Except for faults at short distances from a terminal, it can be concluded that the distributed method yields results which, for all practical purposes, are comparable to the lumped method. As indicated by the preceding equations, the discrepancy between the two methods will depend on the ratio of the structure grounding resistance to the impedance of the section of ground wire between two structures. In most cases this discrepancy will be in the order of 10%. Figure 6.6 shows the current in the transmission line ground wire for a phase-to-ground fault at various distances from the left (source) terminal. The solid-line curves correspond to the results as computed by the lumped method, while the broken lines are for the distributed method results. This figure clearly illustrates the very good agreement between results from both methods. Except for a fault at a very short distance from the terminal, the differences in the results, already very small for a fault at the sixth structure from the terminal, are barely noticeable when the fault occurs beyond structure 14. Finally, it should be noted that because of the electromagnetic coupling between the phase and ground wire, the ground wire current on the left side of the faulted structure does not drop to zero when the fault occurs at a remote structure, but remains at a value equal to the product of the fault current by the coupling factor (Zm/Zg). Similar curves are also given in Volume 2. These curves are based on the distributed parameter method. 6.4 VARYING LINE PARAMETERS Although the overhead conductor impedances are practically constant along a transmission line, the ground resistance varies from one structure to the other, making it unrealistic to assume a constant structure ground resistance along most transmission lines. The methods presented in Section 6.3 are based on this assumption and, although the uniform structure ground resistance value used is a mean value, the results are generally inaccurate when the faulted structure or an adjacent structure has a ground resistance significantly different from this average value. To improve the accuracy of the calculations it is important to represent the faulted structure and its adjacent structures with resistances close to the real values. The remaining structures are then represented by actual values or a mean value. In this case, the previous methods can not be used and new approaches must be developed to satisfy the modified requirements. In this section, a new calculation method will be described. This method lends itself to digital computer processing and therefore was used as the basis for the development of the computer program PATHS described in Appendix D. 6-12 The loop and node equations can be used to solve for the unknown voltages and currents of the equivalent circuit. In solving the equations, a stage is arrived at, where the equations contain the terminal voltages of the span, the branch currents and the circuit impedances. These equations can be written in matrix form as follows: [Uk+1] = [Sk][Uk] ; k = 1 , n-2 (6-36) where a, k /g j k I a I [Uk] (6-37) g5k „ and 1 0 z 0 1 Z 0 0 0 1/Rk [sk] -Z oa -z om 1 Z om om og (6-38) 0 /R. - (1+Z /R.) k og k If there are n spans between the feeding terminal and the faulted structure, Equation 6-36 represents a series of n-2 equations which permits the computation of the currents and voltages in span k+1 from the corresponding values in span k and vice-versa. The equations for span 1 and n-1 are determined from the boundary conditions at the fault location and at span n-1. Based on Figure 6.8, the following can be written: R (I -I J e a g, 1 Vi [U I ,) R e (I a -I g, 1 3 I a I g, ,1 (6-39) a I, 1 and _ V V I I — a, n g>n a g,n . V R I a, n .(I .-I ) n-1 g ,n-1 g,n a I _ g»n 6-14 J (6-40) FauIt at Tower #2 Lumped Parameter Method Distributed Parameter Method Fault at Tower #6 Fault at Tower #12 Pault at Tower #32 "Trapped1 Current STRUCTURE NUMBER Figure 6.6 Currents in Ground Wire Other methods have been developed and are also used. The method described by Sebo in 1969 [6] is an accurate approach which permits varying structure resistances along a transmission line. More recent works have proposed alternative methods using different computation techniques [7-10]. Essentially, all methods are based on the same initial equations. The differences are caused by the methodology and algorithm used to perform the necessary calculations. A brief description of Sebo's method follows. 6.4.1 SEBO'S METHOD In Sebo's method, the section of transmission line between two consecutive structures is represented as a six-terminal network. This six-terminal network is shown in Figure 6.7(b). It consists exclusively of zero sequence self-impedances and is equivalent to the real circuit of Figure 6.7(a). This last circuit was originally proposed by Clarke [11] in order to retain the identity of the earth return path. It is a zero sequence equivalent circuit, constructed on a per phase basis for a three phase transmission line between two grounded structures solidly bonded to the overhead ground wire. 6-13 Equations 6-36, 6-39 and 6-40 result in a chain multiplication of matrices of the type shown by relation 6-38. This chain multiplication can be written as follows: [U ] = [S ] [S ,] n n-1 n-2 [S33 [S23 [S^ [U^ (6-41) Matrix Equation 6-41 results in three linear equations with five unknown quantities, i.e., ^a,ns Ia> *g,l> ^g,n-l anc* ^g,n* One additional equation is obtained from the terminal span loop, that is: (6-42) I a (Z a,n + Rj x, - I g,n (Z ' g,n + R n-1 + Rn) x, + I g ,n-1. R n-1. = 0 -x. Ol > DEFINITION OF SYMBOLS Zoa: Self Impedance of Phase a, Span k Zog: Self Impedance of Ground Wire, Span k : Variable Related to Span k Zoe: Earth Return Impedance, Span k Vxk: Voltage of Conductor x at Span k Zom: Mutual Impedance, Span k Ixk: Current in Conductor x of Span k Zx = Zox-Zoe, Span k (Conductor x) : Tower Ground Resistance, Span k (b) EQUIVALENT CIRCUIT (a) REAL CIRCUIT Figure 6.7 Circuit of One Span of a Line The last equation needed to determine all the unknown quantities is obtained by noting that the sum of the currents in the terminal and transmission structure grounding systems must be equal to zero. In some cases, the total fault current Ia is estimated from previous short-circuit studies or has already been measured. Therefore, no additional equation is required and the unknowns are determined with respect to Ia assumed as a reference 1.0 p.u. value. It should be noted that the chain multiplication shown in Equation 6-41 may lead to serious round-off errors during the computations, when the number of spans is large and/or where the elements of the matrices vary between wide limits. This is similar to what can happen to the single-sided elimination method described in the section following. 6-15 I Figure 6.8 I a a Ground Fault at Span n from the Source Terminal 6.4.2 THE SINGLE AND DOUBLE-SIDED ELIMINATION METHODS The analyses presented in this section are an extension of the work originally published in [9]. Essentially the equations are updated to cover the case of a three phase circuit with mutual coupling between phase conductors. Two methods are described; the single-sided elimination and the double-side elimination methods. The first method leads to equations which express the current in a span as a function of the currents in the preceding (or following) span. This procedure is particularly vulnerable to round-off errors. In contrast, the double-sided method although not immune to round-off errors, is significantly less prone to such errors. This is because the current in a span is determined as a function of the currents on both sides of that span, leading to major differences in the structure of the final equations and computational algorithms used to find the solutions. The problem being analyzed is shown in Figure 6.9. A fault occurs on structure 0 of a transmission line. The fault is either single, double or three phase to ground and the transmission line is a single or double-circuit line. The phase conductors may be transposed or untransposed. There are n and m spans between the faulted structure and the left and right terminals respectively. There is one overhead ground wire (or equivalent neutral conductor) in the transmission line. The parameters of the network shown in Figure 6.9 are defined as follows. In order to improve the readability of the equations, the superscripts used to designate the circuit number (1 and 2 for a double-circuit) and the terminal side (l for left side and r for right side) are dropped except when necessary. Self impedance of phase conductors of the circuit in span k (ohms). Zpk ZXyk Mutual impedance between conductors x and y in span k (ohms). Structure ground impedance at extremity of span (ohms). 6-16 Zfx Impedance of fault path for conductor x (ohms). Zp Equivalent source impedance at terminal (ohms). Zg Equivalent ground wire impedance at terminal (ohms) Rt = rn Ground impedance of terminal (ohms). Re = r0 Ground impedance of faulted structure (ohms). U Transmission Line voltage (volts, phase-to-phase). fi Number of phase conductors (3 for a single circuit, 6 for a double circuit). Equivalent Source impedance Left Terminal Figure 6.9 Phase Conductors Fau1 ted Structure No. 0 Right Terminal Equivalent Network - General Case The data given above are used to determine the currents in the loops shown in Figure 6.10. Loop 0-x is defined by phase conductor x, the ground wire and the fault impedance. Loops 1 to n (or m) are defined by the ground wire and the structure impedance at the extremities of each span. Loops 1 (left and right) have impedance Re in common, while loop n (or m) has the terminal ground impedance as one of its branches. For simplicity, it is assumed that the mutual impedances between phase conductors are all equal and that the mutual impedances between the ground wire and any phase conductor are also equal, i.e.: = ZXyk ; where x or y represent the ground wire. Zuk = ZXyk 5 where x and y represent a phase conductor. This simplification will be removed in the next section. 6-17 Phase J + 1 Phase J j+1>n / “pk MMa-y-y/------ |-AAA/W|------ ---|-AA/WV------- AAAAAry-- ■/(j+i>iCL ej+1 (rv)Zf. X/ X^-i X ------- ^SAA/Wj-------- -jAAAA/V—AAAr-|-------- zgnX "kX'ij ^X^-1 X i- n-1 Ground wire Figure 6.10 zgl fg k-i Fundamental Loops of the Network The following equations govern the current in loops 1, k = 2, n-1 and n: V SF. + T.I. - r. I 11 j = 1 J-J 1 Z + R (I c s - Ij = 0 1 ■ y Er rk-1Ik-1 + TkIk " rkIk+1 Lj=l JJ H nSr Lj = 1 r ,1 (6-43) = 0 , + T I n-1 n-1 0 n n where S, k = Z . gk k mk k-1 There are m pk (6-44) k additional equations for each loop of type 0: y H - Z . n Ef: - E sk\ + GF. + Zf;l! = E. ; j=1 J "f i i k=1 i for i = 1 ,y (6-45) where n ^pk G = ^mk ^uk^ (6-46) n H = E k=l (zgk " Zmk + Zuk^ a 6-18 and Ej is the source voltage of phase i. The module of Ej is U and its angle will be 0°, 120° or 240° depending on whether conductor i represents phase a, b or c respectively. The currents Ie and 1; are defined as follows: I e = i£ + ir 1 1 If = F? ♦ I ! Fr. ! (6-47) ; (6-48) for i = 1, y The Single-Sided Elimination Method The form of Equations 6-43 suggests that the loop currents 1^ (k = 2, n) can be expressed as a function of the total left terminal fault contribution , the faulted structure current Ie and the preceding loop current Ij.^. This is easily proved as follows: First the equation related to loop k-1 is rewritten as: (6-49) Xk = Vk-2 + Vk-I + CkF* where Fil = M ZFj j=l 1 Ak ' ' rk-2/rk-1 (6-50) Bk = Tk-l/rk-l Ck ' ‘ Sk-1/rk-l Then, from the equation related to loop 1 the current I2 is written as: I2 (6-5T a2I£ + a2I1 + 62Ie where “2 = C2 = -Sl/r1 °2 = B2 + A2 = (TrRe)/rl = (Zpl + rl)/rl (6-52) 62 = -A2 = Vrl At this point, it can be proved by induction that Equation 6-51 is also valid for k = 3, 4,—,n. Based on the initial values given below, this general relation is: Ik = (6-53) + Vk-I + 6kIe where “k ' Akak-2 * Bkak-1 + C 0k ' Vk-2 + Bk0k-I (6-54) Sk = Vk-2 * Bk6k-1 6-19 The initial values of cr^, and being: a0 = al = o = a1 = 1 (6-55) 6q = -1 and 6^ = 1 Assuming that the values of I^, Ie and are known, all loop currents can be determined using Equation 6-53. In this equation the current in loop k is computed following the calculation of the preceding loop current (k-1). It should be noted here that the value of oq , <rj and <5j can vary, at least theoretically, between wide limits (small to large values). For that reason, the single-sided method is not suitable for computations involving large numbers of spans and/or impedance values differing from each other by several orders of magnitude. A detailed analysis on the sensitivity of the method to variations in the number of spans and impedance values is not of interest to this study. However, the following example will clearly illustrate the round-off errors which can be caused during the computations. When it is assumed that there is no electromagnetic coupling between the phase conductors and the ground wire (Z^ = 0), the sum of a [<+ becomes equal to 1. This is proved as follows. From Equation 6-34, this sum is found to be: “k * °k = Since <*1+ Vv2 anc] + 0k-2> + Ek (V, * Vi’ + Ck are equal to 1 (Equation 6-55) and since: aQ+ Ak + Bk + Ck = ^"rk-2 + Tk-1 " Sk-1^/rk-1 = 1 Then the value of +. will remain constant and equal to 1 for all spans k. Assuming that Zg^ = 1 f], r^ = 0.1 Q and Re = r0 ^ 0 Q, then Table 6.1 gives the computed values of ana as performed by a 32 bit computer (^ 7 digits) and a 10 digit hand calculator. COMPUTER CALCULATOR K ak °k V°k 0 1 2 3 4 5 6 7 8 0. 0. -10. -130. -1560. -18600. -221650. -2641210. -31472880. 1 . 1 . 11 . 131 • 1561 . 18601. 221651. 2641210. 31472848. 1. 1. 1. 1. 1. 1. 1. 0. -32. Table 6.1 ak 0. 0. -10. -130. -1560. -18600. -221650. -2641210. -31472880. ak 1 . 1 . 11 . 131 . 1561. 18601. 221651. 2641211. 31472881. Vak 1. 1. 1. 1. 1. 1. 1. 1. 1. Round-Off Errors In Computations The round-off errors are clearly illustrated in this table. Although it is possible to eliminate this type of problem using suitable techniques, it is preferable to use a different approach which is not subject to such complications. The double-si ded method is one possible approach to this problem. 6-20 The Double-Sided Elimination Method This method follows the same steps as for the single-sided method up to Equation 6-48. However, instead of using Equation 6-49, Equation 6-43 is rewritten to produce the following equation (k = 2, n-1): (6-56) = Vk-1 + BkIk+1 * v* where Ak rk-1/Tk (6-57) Bk = rk-1/Tk Ck = Sk/Tk The and defined by (6-57) are always less than or equal to 1. This is a significant computational advantage which drastically reduces round-off error problems. Equation 6-56 becomes valid for k = 1 and n when the following definitions are made: A] = -Re/(TrRe> (6-58) B1 = C1 = and An “ (6-59) Bn " -Rt/(Tn-Rt> Cn = Sn/(T„-Rt> Through easy but lengthy algebraic transformations it can be shown (see [9] for calculation details) that the currents in loops 2 to n can be expressed by the following equation: (6-60) Ik = V* + VkIn + Ve and the currents in loops 1, n-1 by: (6-61) Ik = ukI£ + vkIn + wkIe where Uk = (ak + 6k\-1)/Ak Vk = ak/Ak (6-62) Wk = 6k6k-l/Ak Ak = 1 " 6kek-1 uk = (nk + ek ak+1)/Ak (6-63) vk = ekak+1/Ak wk = 0k/Ak 6-21 and for k = 2, n: 5k' (6-64) 0k = °k*i5kBk/Ak “k ' ^kV, * Ck)sk/Ak and for k = 1, n-1: £k = Bk/(l"Ak£k-1) 9k = ek-iekAk/Bk (6-65) ^ = (W-,+ Ck)ek/Bk Equations 6-64 and 6-65 are based on the initial values given in Table 6.2. 6. i ai a. i £. 1 0. I ni 0 1 0 0 i 0 n B1 A1 ci i=n+1 A i =n Table 6.2 B n n c i=0 i =1 Initial Values Detailed calculation shows that the 8, <r, e and tj coefficients are positive values which can not exceed 1. The <r and 6 coefficients are also positive and are bounded as follows: a. < i 0. < i R .+Z n-1 gn V "gi This is also another major reason why the method is less subject to round-off errors. Equations 6-60 and 6-61 allow the calculation of the loop currents as a function of the total fault contribution from the terminal (F^), the terminal ground system current (In) and the current in the faulted structure (Ie). These currents must be determined before one can proceed with the loop current calculations. Replacing 1^ in Equation 6-45 by its expression on the right-hand side of Equations 6-60 and 6-61, yields: j=1 IF* - fcl (SjV,A I n E S. v. )I k k n (6-66(a)) - (SjWj 6-22 ?Bk«k) p-r _ _£ Zj-.r . - E. fI I i where i = 1, m (m is the number of phase conductors) and Fjr conductor i of the right terminal. is the current in phase A similar equation (6-66(b)) is applicable for the right terminal: lur cr r (H -S,U| PK) E 2 + - (S^ + j=1 ^ (6-66(b)) eT z^.F. = (sW + EsX1* fi i 1 Therefore, Equations 6-66(a) and 6-66(b) represent 2m equations with 2m + 3 unknowns. Three more equations are needed. Two equations are obtained from Equation 6-43 applied to the terminal loops (k = n for left terminal and k = m for right terminal): (S* + r v n U P J E F • + (T ^ j n j=1 n-1 n-1 H £ ..5, £ F1; + (Tr - rr 1Vr Jir - / 1Wr m-1 m-1 *—1 j=1 j m n T _ I = 0 (6-6?(a)) ,V .) I - r , W ,1 - 0 n~1 n-1 n n-1 n-1 e y (Sr + rr 1Ur .) m r m-1 m-1 m (6-67(b)) m-1 m-1 e and the last equation is obtained from Equations 6-47 and 6-61: £ u, 1 Equations n „ r£ . F. + f—1. j >y j=1 r u. 1 6-66, 6-67 m cr . X F- + - , J J=1 and 6-68 £T£ , r r v.I + v,I + 1 n 1 n ,..£ . .r (W. + W, 1 1 - 1)1 e = (6-68) 0 provide the solutions for the unknowns Fj, ln and Ie. The double-si ded elimination method is well suited for computer-based solutions. It provides fast and accurate results without a large problem with undesirable effects, such as round-off errors. The computer program PATHS developed within the scope of this study is not based on this method but rather on the following more general form of the method. This choice gave the program increased capabilities and more flexibility to implement future extensions and enhancements. In addition, because the equations developed are more general, they can be used, after minor modifications, in a variety of other but still related problems such as electromagnetic induction and substation grounding analyses. 6.4.3 THE GENERALIZED DOUBLE-SIDED ELIMINATION METHOD The power system network being analyzed is shown in Figure 6.9. The parameters shown in this figure are as described in the preceding section. The phase conductors of the network may be considered as a number of elementary loops consisting of short conductor segments between pairs of nodes, each grounded via a fictitious very high impedance ground. When this concept is adopted, the fundamental loops of the network may be selected as shown in Figure 6.11. In this figure, the loops related to the phase conductors correspond to the transmission line spans. Hence, the nodes will be located at each transmission line chain of insulators. The concept introduced in the preceding paragraph is not unreasonable. The high impedances to ground of the nodes are equivalent to the per span capacitances of the phase conductors. These capacitances have not been neglected in the derivation of the final solution. However, as can be observed from Equation 6-7Q(b), the effect of neglecting the capacitances can easily be accounted for in the final equations. 6-23 Loop 0-y r =R Loop k Figure 6.11 Fundamental Loops of the Network The equations which follow are established using the left side of the network (left terminal). Except when necessary to avoid ambiguity, the superscript l (for left) has been omitted from the variables. It is obvious that similar equations can be written for the right-side loops of the network. The following equations apply for loops k = 2 to n-1 of each phase conductor i (i = 1, M): y+i , 1 1 Xk " aikIk-1 T T “ik^l = 0 (6-69) j=i where sijk ■ <2ijk * Rt>/Tik ; for I - 1,y and j = 1 ,y sijk ' <2igk * Rt>/T,k ; for i = 1 ,1J and j = p + 1 = g a., ik = ck-,/Tlk bik = c.k /T.. ik T., ik = c, . + Z., + c k-1 ^ik Lk (6-70(a)) When the line capacitance is neglected, i.e •> ck is assumed to be a very high impedance (c(< = oo), (6-70(a)) becomes: a.k =0.5 bfk =0.5 (6-70(b)) 6-Zk An equation similar to 6-69 applies for the ground wire i = /* + 1 = g Y' A-! j=1 s .. i/ + i g _ a..Ir, . gjk k i k k-1 bikIk+1 = 0 (6-71) where s_.i = Z ../I . gjk gjk gk ; for j = 1 ,p J ’M aik = a9k " rk-1/Tgk (6-72) ^ik kgk rk'/^gk T . = r, . + Z . + r, gk k-1 gk k Equations 6-69 and 6-71 can be condensed into the following matrix equation: (6-73) csk]CFk] = CAk]CFk-i] * [VCW where the matrices are diagonal, column or square and are defined as follows: Span Current Matrices (unknowns) 1 Ci H 1 k-1 k [Fk] " ; ' [Fk-i] = ipk CFk+1] " C, 1 . i9k. I9 k-1_ Coefficient Matrices '12k s K 21k « " v v [Sk] N ylk ’gi k- gyk 6-25 0-- 0 0 a2k 0 0 \ c/y \ -0 a 0 0 0 Vik agk b.. 0ik 0 vb 2k N [Bk] -0 o -0 I I o \ 1 n0 -0 b o yk b gk Matrix Equation 6-73, when applied to the left and right sides of the network, provides (m + 1) (n + m - 4) simultaneous linear equations to determine the (m + 1) (n + m) unknown span currents. The 4(m + 1) equations still required to completely solve the system are obtained from the boundary conditions at the left and right terminals and on both sides of the faulted structure. In the following, only the left side boundary equations are established. Left Terminal The following matrix equation holds at the last loops (terminal loop) of the transmission line: [S ] [F ] = [A ] [F ,] + [£] n n n n-i (6-74) where Ign 12n21n i CS 3 n i i 8 v \ X \ X S 5g1n- V • gyn 6-26 1 a, 1n 0?2 2n [An] EE] \ i x 0 —0 a gn for i = 1, y: s. . = (Z.. + RJ/T. i jn j in t in for j = 1, s . = (Z . + R )/T gm gin t' gn = a. cn~1,/T.in in = r ,/T a n-1 gn gn e. = E./T. i in T. in = c T = r gn y + 1, j * i (6-75) , + Z. + R. n-l in t , + Z + R n-1 gn t Side of Fault Location The following matrix equation can be written for the first loop of the transmission line: CF1 ] = [B.|] [I2] - CH] (6-76) [F0] where 1 s 121- S211 x igi \ \ [S^ s gi 1 ygi ’gyn b11h °b21 [B1 ] 0 0---------------------- o bg1 6-27 ‘11 12- ig 22 21 [H] ■hgy gi- and, for i = 1, n: for j = 1, y + 1, surzrn/T-n SgI1 99 j i i Zgi1^Tg1 bi1 = C1/Ti1 b . = r./T gi 1 gi Hn = <zh + V/Tn (6-77) hgg - Re/Tg1 h.. = R /!., iJ e 11 H . = R /T , gj e gi T., = Z.,+ c, 11 i1 1 T 1 gi = z , + r, gi i The fault path current matrix [F0] has been introduced in Equation 6-76. This matrix is defined as follows: ■1£ -1 .21 '1 [F0] = Matrix [F0] introduces equations. y£ 1 gj> m + + I1r I1 I2r I1 > 1 = o [Fp [pP (6-78) 1?" + 1 unknowns and matrix Equation 6-78 generates m + 1 linear There are now as many equations as there are unknowns. Therefore, a solution can be obtained by inversion of the complete matrix system or using other algorithms suitable for solving a set a simultaneous linear equations. Obviously, this approach is reasonable only when the number of spans in the network is relatively small. In practice, this is seldom the case. Therefore, it is necessary to reduce the size of the system to a smaller number of simultaneous equations. 6-28 The following double-sided matrix elimination method will allow each loop current to be expressed as a function of the fault path current [F0]. Elimination Procedure Starting at the Terminal Equation (6-74) can be written as: [F n ] = [an ] [I n ,] + [A n ] (6-79) -1 where [a ] = [S ]“1 [A ] n n n (6-80) [V = [V~' [E] Rewriting Equation 6-73 and taking into account Equations 6-79 and 6-80 results in the following recursive formula valid for k = 2, n: [Fk] = Cak] [Fk_1] + [Ak] (6-81) where Ca^ = [N,]-' [A,] :xk: = cn,]-' CBk: nk+1: [V - tsk] - [Bk] [tW (6-82) Elimination Procedure Starting at the Fault Location Similarly Equation 6-76 is written as: [F^ = [F2] + [F0] (6“83) where Ir}^ = [S^-1 [B1 ] [a1 ] = -[S^-1 (6-84) CH] By taking into account Equations 6-83 and 6-84, Equation 6-73 can be rewritten as a recursive formula valid for k = 1, n-1: CFk]" [V [,W + Cak] ^ (6‘85) where “k3 Cak] = [Mk]"1 CHk] = [Sk] [Ak] [ak_1] (6-86) - [Ak] [Vl] 6-29 Double-Sided Elimination. At a given loop i = 1, n-1 , matrix Equations 6-81 and 6-86 become: ] tr.] . [A.+1: [f.: - cn,: [Fi+|: + [a.] [f0: Multiplying the first of these two equations by r?j and adding both equations yields: [F.] = [U.] + [V.] [Fq] ; for i = 1, n-1 (6-8?) where [U.] = [W.r1 Cn.] [V.] = CW.]'1 [c.] [W.] = [u ] - En.] [A.+1] (6-88) [a.+1] ; [u] is the unit matrix Matrix Equation 6-87 gives the currents in span i based on the currents [F0] circulating in the fault paths. These currents are determined from Equations 6-78 and 6-87 applied to span 1 on the left and right sides of the fault. The following expression is obtained: [Fq] = [F^] + [Fj] = cuj] + [uf] + { [V^] + [vj] 1 tFo] which reduces to: CF0] = | [u] - [V^] - [v!|] I"1 | cu|] + [u!|] | (6-89) where [u] is the unit matrix [VJ and [UJ are determined from Equation 6-88. 6.4.4 PRACTICAL CONSIDERATIONS Ideally, transmission line ground fault computations should be made using the actual conductor and structure ground impedance values. The conductor self and mutual impedances have been discussed in Chapter 2. Although soil characteristics influence these impedances, the use of an average soil resistivity yields impedance values only a few percent different from those which could be measured directly. Structure Resistance Values Generally there are large variations in soil structure along the route of a transmission line, and hundreds of transmission line structures. Often, the ground resistances of a large number of these structures are not known. Sometimes, an average ground resistance value can be assigned to a group of structures based on the shape of their footings, their installed grounding system and earth geological characteristics common to the sites where these structures are 6-30 located. As can be seen from the example shown later, two situations can develop depending on which structures are the ones for which there is no data or only an average ground resistance value. When these structures are located at some distance from the faulted structure, even large variations in the average resistance values have little effect on the magnitude or distribution of the fault current. This is clearly illustrated by the results of Figure 6.13. These results are related to the transmission line of Figure 6.12 where a fault occurs at structure 72 from the left terminal. BASE 10 10 10 10 A 1 19 10 10 B 10 10 100 100 c 10 10 10 10 D 10 10 10 10 CASE Figure 6.12 T0WE * RESISTANCE 10 10 10 10 10 10 10 — — 10 10 10 10 10 10 1 10 10 10 10 10 10 10 10 19 1 10 10 10 10 19 1 10 10 10 1 10 (IN OHMS) A Phase-to-Ground Fault on a Transmission Line For example, a factor of 10 change in the resistances of four structures located midway between the left terminal and the faulted structure (R37 = R38 = 100 ^ and R39 = R40 = 1 fi), does not change the currents at or near the faulted structures by more than a fraction of one percent. The same conclusion applies when the resistances of structures 1 and 2 (the structures which are the closest to the left terminal) are changed from 10 to 1 Q and 19 Q, respectively. In this last case however, the current in structure 1 rises from 2.9 A (10 case) to 22.3 A (1 case). In both cases the total fault current remains constant. However, when the faulted structure and/or one of its adjacent structure resistance values are not well known, significant differences are obtained between computed results based on the various assumed values. This situation is also illustrated in Figure 6.13. The preceding discussion indicates that it is important to use accurate ground resistance values for those structures which are closest to the point of fault. Utilities which measure the resistance (or the soil resistivity) of all their transmission structures need only decide whether to use the actual resistances of the remote structures or an average value. When not all the required data are on file, a decision must be made as to whether ground resistance (or resistivity) measurements should be conducted at the site of interest. Fortunately, such predicaments are scarce in practice. When a fault current distribution analysis at a given structure is necessary, that structure is usually either special or in a special area. In either case, the structure may have already been the subject of particular attention and suitable data about the structure and the soil conditions where it is erected will typically be available. 6-31 For example, the structure may be located in a densely populated area and require a specially designed grounding system. Alternately, an incident caused by a ground fault at a structure may require a detailed investigation into the cause of the incident. In such cases, detailed measurements are generally required. Mutual Coupling The mutual impedances between the ground wire and the phase conductors reduce the apparent impedance of the circuit seen from the feeding terminals. Therefore, the value of the total fault current calculated is larger than the value computed neglecting these mutual impedances. The mutual coupling has also another effect which is to force the circulation of "trapped" current in the ground wire which returns directly to the source. This current plays a major role in reducing electromagnetic induction on neighbouring circuits. This effect is commonly designated as the ground wire screening effect. Terminal Structures When the transmission line is relatively short, all structures will carry the fault current to ground in the same manner as for the faulted and adjacent structures. However, in the case of long lines, the terminal structures start collecting currents from earth as if these structures were part of the terminal grounding system. 1600- aicc cc D Q D < LL _) < oHhUO Case Base Total Fault I^ (in Amperes) 2757-667 A 2757.673 B 2757-667 C 2811.870 D 2793-650 Case C (R72 = U2 ; R71 = 19 it) z HIcc h-OD 3 <0S 85 STRUCTURE NUMBER Figure 6.13 Effects of Ground Resistance Values 6-32 90 148 Central Structures For a long transmission line, the structures located in the central area between a terminal and the faulted structure conduct at most, a negligible current to earth. Therefore, these structures need not be modelled in the circuit. This is an important effect, since it permits the analysis of long and short tramsmission lines based on a limited number of structures represented in the circuit. For example, if a maximum of 100 structures per terminal is permitted by a computer program, transmission lines up to approximately 20 miles in length can have all structures included in the model. For longer lines, it is necessary to reduce one or several sections of the transmission line into equivalent Pi sections. For example, all structures can be included for 10 miles at each extremity of the transmission line and the remaining central portion of the line can be replaced by one equivalent single span of appropriate length. REFERENCES 1 - Janos Endreyni, "Analysis of Transmission Tower Potential During Ground Faults", IEEE Transactions on Power Apparatus and Systems, Vol. PAS-86, No. 10, October 1967. 2 - R. Rudemberg, "Transient Performance of Electric Power Systems", McGraw-Hill Book Company, 1950 (book). 3 - F. Desieno, P. Marchenko, G. S. Vassel, "General Equations for Fault Currents in Transmission Line Ground Wires", IEEE Transactions, Vol. PAS-89, No. 8, November/December 1970, pp. 1891-1900. 4 - R. Verma, D. Mukhedkar, "Ground Fault Current Distribution in Sub-Station Towers and Ground Wire", IEEE Transactions, Vol. PAS-98, No. 3, May/June, 1979, pp. 724-730. 5 - A. J. Pesonen, "Effects of Shield Wires on the Potential Rise of HV Stations", Offprint from SAHKO-Electricity in Finland, 53, 1980, No. 10, pp. 305-308. 6 - S. A. Sebo, "Zero Sequence Current Distribution Along Transmission Lines", IEEE Transactions, Vol. PAS-88, No. 6, June 1969. 7 - C. Dubanton, G. Gramd, "Influence of the Location of the Fault on the Screening Effect of Earth Wires", CIGRE Paper 36-01, August 1974. 8 - F. Dawalibi, D. Mukhedkar, "Ground Fault Current Distribution in Power Systems - The Necessary Link", IEEE Paper AF77 754-5, Summer Meeting, Mexico 1977. 9 - F. Dawalibi, "Ground Fault Current Distribution Between Soil and Neutral Conductors", IEEE Transactions, Vol. PAS-99, March/April 1980, No. 2, pp. 452-461. 10- F. Dawalibi, D. Bensted, D. Mukhedkar, "Soil Effects on Ground Fault Currents", IEEE Transactions, Vol. PAS-100, No. 7, July 1981, pp. 3442-3450. 11 - E. Clarke, "Circuit Analysis of AC Power Systems", Wiley, New York 1956. 6-33 LIGHTNING PERFORIVIANCE OF TRANSIVIISSION LINE STRUCTURES 7.1 GENERAL Literature on lightning phenomena and related effects is voluminous. Even when limited to the subject of power system engineering, the list of research work published in the last 50 years remains impressively extensive. A meaningful review of the important literature which has contributed to a better understanding of transmission line performance during lightning activities involves hundreds of papers and reports published by distinguished researchers throughout the world. This formidable task is beyond the scope of this chapter which concentrates on the response of transmission line structures to direct lightning strokes. Furthermore, no attempt is made to examine the problem in a broad power system engineering perspective which would necessarily involve a probabilistic approach in order to reflect the statistical nature of lightning and its effects on power systems [1,4]. This chapter reviews analytical methods and computation techniques presently available to determine the performance of a transmission line structure struck by lightning. 7.2 EFFECTS OF LIGHTNING ON POWER LINES Among the many events which can lead to a power system outage or equipment damage, lightning is certainly the major identifiable cause [2]. This has been so since the earliest days of the electric power industry. There are an estimated 1000 thunderstorms at any moment in time throughout the world. The most visible manifestation of a thunderstorm is the thundercloud, a strongly convective cumulo-nimbus cloud. The phenomena which take place inside this cloud are not well known because of the extremely violent internal turbulence. Moreover, the mechanism by which thunderclouds become electrified is still the subject of controversy [1]. The lightning discharge mechanism has been the subject of extensive research for many years however, and is now well understood. The following is a very brief summary of the lightning discharge process. A comprehensive description of this process can be found in [1]. 7.2.1 THE LIGHTNING FLASH The lower portion of a thundercloud is usually negatively charged while its upper part carries positive charges. Thus, most cloud discharges to ground are negative discharges. Occasionally, positive discharges are produced by clouds with a small positive charge at their base. 7-1 The leader stroke is always initiated at a point of maximum electric field strength. This point is generally located at the base of the thundercloud, initiating a downward leader. Less frequently, this point is located at the top of a tall building or structure or elevated ground, resulting in an upward leader, often termed an upward streamer. Frequently, this upward streamer develops as a downward leader approaches the earth. In the initial stage of the formation of a negative lightning stroke, the downward leader starts at the negatively charged base of the cloud and proceeds towards ground in successive steps, separated by time intervals of 40 to 100 ms (stepped leader). Each step follows a seemingly random path and is believed to be practically independent of the nature, configuration and number of the various grounded objects below. However, as the electric charges in the cloud follow the leader stroke at an average speed of 100 km/s, the electric field on the surface of grounded structures increases rapidly until a critical electric field strength of approximately 500 kV/m is reached, causing the initiation of upward streamers from the most prominent structures. The first discharge occurs when one of these streamers reaches the tip of the downward leader. Following the first discharge or return stroke, there may be one or several successive discharges of various magnitude and duration. The distance from ground at which the upward streamer is initiated is called the "striking distance" and plays a major role in the design of shield wires for the protection of transmission lines against lightning outages. The preceding discussion of the leader stroke mechanism suggests that ground resistance may not be a factor in determining the probability of a lightning stroke to a particular structure. This conclusion however, is challenged by recent statistical and laboratory observations which confirm the common belief that lightning is attracted by low ground resistance structures [3]. The attract!vennes of good grounds to lightning can be explained by the process taking place during the last stage of the stepped leader progression towards the earth. As the distance to earth decreases, several streamers are initiated from various locations. However, the major portion of the discharge will occur along the path of lowest impedance. This explains the occurence of simultaneous strokes at different locations. It is important to note that the lightning susceptibility of low resistance grounds should not be interpreted as a reason to seek high resistance transmission line structure designs in an effort to reduce the number of strokes on the line. Regardless of structure resistance, the transmission line remains the preferred target for a lightning discharge and the transmission line structure will dissipate a direct lightning stroke more easily if it has a low resistance value. 7.2.2 THE BACK FLASHOVER MECHANISM Fifty years ago, it was generally believed that protection of transmission lines from direct strokes was impossible. Now, all but a few transmission lines are designed to withstand a large proportion of direct lightning discharges. Since the importance of service continuity is constantly increasing, research efforts are being directed towards the development of improved protection equipment and more accurate methods of predicting the lightning performance of transmission lines, particularly with regard to the response of structures to direct lightning strokes. There are three mechanisms by which insulation flashover results from a lightning stroke: a - Electromagnetic induction, b - Direct stroke or shielding failure, c - Back flashover. 7-2 Transient voltages are electromagnetically induced in power lines as a result of lightning strokes occuring in the vicinity of the lines. These induced voltages are caused by the very rapid collapse of the electric field due to the lightning current discharge. This type of induced voltage, although the cause of numerous problems on distribution systems, is generally harmless to tranmission lines rated 100 kV or more. This source of distribution line insulation failure is not examined further in this study. The reader will find a wealth of useful information on this subject and references in [1]. Even when a transmission line is equiped with shield wires optimally located to intercept lightning strokes, the line may, occasionally, be subject to a direct stroke. This possibility is explained by the previously defined "striking distance", which is the basis for various electrogeometric theories developed to explain outages due to the shielding-failure flashover mechanism. These concepts will not be developed here and reference is again made to [1] as a comprehensive source of additional information. This chapter is devoted to the back flashover mechanism, defined as the insulation failure due to excessive rise in potential of the transmission line structure when struck directly by lightning. The case of a lightning strike to the overhead shield wire at midspan can be analyzed by essentially the same methodology. Discharge of lightning current into earth via a transmission line structure may raise the potential of the structure with respect to remote ground to several times the transmission line voltage. If this value exceeds the dielectric strength of the transmission line insulators, a back flashover occurs, soon followed by a power frequency arc fault along the ionized air path created by the lightning arc. This mechanism is illustrated by Figure 7.1. IDEALIZED LIGHTNING SOURCE GROUND WIRE stress FOOTING GROUNDING REMOTE GROUND (a) Figure DIRECT LIGHTNING STROKE 7.1 (b) EQUIVALENT SIMPLIFIED CIRCUIT A Typical Transmission Line Back Flashover The majority of lightning outages on shielded transmission lines are the result of back fl as hovers, since only a small porti on of the strokes directly strike the phase conductors of an effectively shielded line. 7-3 At the instant of impact of a lightning stroke on a transmission line structure, current waves originate at the point of impact and travel along the various paths at velocities between 0.2 and 0.9 the speed of light. Current waves move along the overhead ground wires towards other structures and a current wave flows to ground through the structure and its grounding system, including the structure footing and counterpoises, if any. Upon arrival at a terminal (local grounding system) or junction points, (adjacent structures) the waves are partially reflected and return to the point of impact where they interfere with the continuing initial current discharge. This process is depicted in Figure 7.2. Generally, successive reflections up and down the affected structure and from adjacent structures reduce the voltage buildup across the insulators well before the lightning discharge process is completed. Hence, the time of maximum likelihood of an insulation breakdown is generally at the very beginning of the lightning discharge for fast front waves, and a few microseconds after the initial stroke for slower front waves. LIGHTNING__ STROKE WAVE CONCENTRATED GROUNDING SYSTEM i GROUND WIRE PHASE CONDUCTOR INDUCED. WAVE INSULATOR STRING Figure 7.2 ORIGINAL / WAVE COUNTERPOISE REFLECTED WAVE Travelling Waves After a Lightning Stroke to a Structure The usual criterion for the occurence of a back flashover is that the calculated overvoltage curve intersects the insulator volt-time curve. More sophisticated criteria could be used to account for variations in the impulse strength with waveshape [4]. However, a reliable and realistic flashover prediction is dependent on the accuracy of the computation of the stress voltage applied across the insulators. This accuracy is essentially dependent on the reference analytical model and assumptions used to simulate the back flashover mechanism. It is also heavily influenced by the selected values of parameters which singnificantly affect the physical process. Over the past fifty years, several analytical models of varying degrees of complexity have been proposed for back flashover computations [5,7,11-15]. As yet, none of these methods have exhibited an obvious superiority over the others. The reasons for this will become evident as we proceed through this chapter. Nevertheless, it should be noted that most of the researchers working on these analytical models have reported reasonable agreement between their computations and measurements. Furthermore, as will be illustrated, different analytical models will usually lead to similar results if the initial values of the variables are judiciously chosen [5]. Clearly, it is a situation where analytical and computational difficulties impose too many simplifying assumptions which, although appropriate in some cases, are not always justifiable. Furthermore, because field measurements are equally complex and difficult to interpret, the experimental results have not only failed to answer the controversial issues, but have in some cases, introduced additional confusion to this intricate subject. 7.3 ANALYTICAL MODELS FOR BACK FLASHOVER COMPUTATIONS Figure 7.2 shows the passive elements of the lightning current path which, if present, will have a major influence on the lightning performance of the transmission line structure: - Overhead ground wires. Phase conductors. Phase insulators. Transmission line structures. Concentrated grounding systems. Counterpoises. In addition, characteristics of the current source, or lightning stroke channel, such as current magnitude and wave shape can significantly affect the lightning performance of the structure. The accuracy and complexity of the analytical model used in the computations will depend not only on the elements and source parameters retained in the simulation, but also on their representation in the model. This section describes several models used by design engineers. 7.3.1 SIMPLIFIED MODELS The solution of electrical transient problems leads to differential equations which, unfortunately, very quickly become untractable even with only moderate increases in the complexity of the electrical circuit. This difficulty is somewhat alleviated by the use of the Laplace transform which converts the differential equations into equivalent algebraic equations, similar to the ones obtained by a steady state analysis of the problem [6]. However, the inverse Laplace transformation, required to obtain the final solution in the time domain, generally cannot be obtained analytically. Practically, however, there are several techniques, most of them suitable for computerized applications, which provide the inverse transformation with acceptable accuracy. Various authors have proposed simple electric circuits consisting of lumped elements as a model for determining the potential at the top of the structure [7-11]. The most widely used circuit is based on the so-called AIEE method [12]. Three typical models are shown in Figure 7.3. Because of the symmetry of the problem, the transmission line is shown folded back on itself at the structure that was struck by lightning. Hence all the original circuit parameters, except those related to the structure which was struck, are halved. 7-5 GROUND WIRE (l-A)s(t) PHASE CONDUCTOR (a) Figure 7.3 (b) AIEE METHOD (c) Simplified Models The Inductance/Resistance Method The simplest approach is to neglect all parameters except the structure ground resistance, R. If i(t) represents the portion of the stroke current flowing in the structure at time t, then the structure top potential with respect to remote ground is simply: e(t) = Ri(t) For very high values approximation gives leads to increasingly resistances of 10 to (7-l(a)) of structure resistance (say about 100 ohms and more), this very rough a reasonable order of magnitude approximation of the top potential. It optimistic values as the resistance decreases to more typical structure 20 ohms. The results are considerably improved if the transmission line structure is represented by an additional lumped inductance L, as shown in Figure 7.3(a). In this case the structure top potential is: e(t) = Ri (t) + L -—^1 (7“ 1 (b)) The structure current i(t) is a function of the stroke current s(t), the lightning channel circuit and other circuit paths. Solutions are provided in [7] when the stroke current is represented by a double-ramp wave. For arbitrary current waves, the solution may be obtained as follows. First, a unit step stroke current u(t) (i.e., u=0 for t<0 and u=l for t>0) is assumed to strike the structure top. Based on the parameters of the circuit model used, the current in the structure is then determined. Let this current be g(t). Note that g(t) is often referred to 7-6 as the indicial response of a system to the unit step stimulus u(t). For example, in Figure 7.3(b), the fraction of the stroke current flowing in the overhead ground wires is estimated as a constant fraction (1-A) of the total current. Therefore: g(t) = Au(t) (7-2(a)) Generally, A is a complex function of time and of several circuit parameters. From Figure 7.3(a), the following equation is obtained: Z g(t) = T - [ 1-e'at ] u(t) g (7-2(b)) 2r 7 where 2R+Z a g ~2T Next, the superposition technique is used to derive the current i(t) of the structure following an arbitrary stroke s(t). The stroke s(t) is approximated as the sum of a series of step functions of appropriate magnitude suitably displaced in time (see Figure 7.4(a)). The total structure current i(t) is then expressed as the sum of the individual currents computed using Equation 7-2, for each previously defined step function: i(t) = s (0)g(t) + [A. g(t-t. )] k=1 where K (7-3) K may be selected as: xk. k ■jViLLlilki flT V, - tk s (t) CURRENT WAVE s(t) s(t)=at (t)=[at+B]u(t-T) 0 tit k k+1 TIME (a) ARBITRARY WAVE Figure 7.4 (b) DOUBLE-RAMP WAVE Approximation of a Current Surge 7-7 Alternatively, a closed analytical form solution may be obtained from Duhamel's integral: i (t) = s (0)g(t) +J~ -j- g(t-T)dx (7_/*) A close examination of Equation 7-4 reveals that it is the limit of Equation 7-3 when the interval At between successive times, (t^, t^i) decreases to zero (see also the definition of Xk). For the double-ramp stroke current illustrated in Figure 7.4(a) and the circuit of Figure 7.3(c), K on cel [7] obtained relatively simple expressions for the current i(t) for times less than twice the travel time of a reflected current wave returning from the adjacent transmission line structure. Further sophistication can be introduced by including the structure and grounding system capacitances to ground. This is usually of no practical interest since the extremely low time constant, RC, results in negligible capacitance effects for normal stroke current rise times. The AIEE Method The AIEE method illustrated in Figure 7.3(b) assumes that the structure top potential is equal to the voltage drop across the structure ground resistance R plus the voltage determined from the structure self-inductance, assumed to be a constant 20 mH. The potential across the insulator strings is obtained from the structure top potential, decreased by an amount equal to the electrostatic voltage induced in the phase conductor by the overhead ground wires. This method was translated into charts and used extensively for many years [12]. 7.3.2 TRAVELLING WAVE MODELS Although sometimes accurate, the simplified models often fail dramatically in predicting the lightning tripout rates of some EHV lines. Apparently, the tall transmission line structures and their relatively low ground resistances cause the voltage drop across the structure to become a major fraction of the top potential. Thus, preference is given to the use of the surge impedance of a structure rather than the lumped inductance. This approach is further justified through better accuracy in the calculation of steep-front current wave effects. It is also consistent with other traditional methods used in lightning performance calculations which account for the surge impedance of various lightning path components, such as phase conductors, ground wires and counterpoises. Finally, it is more suited for use with travelling wave models and it is in reasonable agreement with advanced concepts of electromagnetic field theories [5,13-15]. A Simple Equation If the transmission line structure and lightning stroke path surge impedances are assumed equal to Zg and all travelling wave reflections except the first one at the base of the affected structure are neglected (see Figure 7.5), the potential of the top of the structure becomes [16]: e(t) Zsi(t) [u(t) + R -Z , irsr U<t-2T>] 7-8 (7-5) where Rs is the structure ground resistance and T is the time of travel of the current i(t) from the top to the bottom of the structure. If the transmission line structure is represented as a lossless transmission line, the surge impedance Zs is: Z (7-6) =yil/c' where l and c are the per unit length values of distributed inductance and capacitance of the structure. In practice, the surge impedance of transmission line structures varies between 60 to 180 ohms. The lightning path surge impedance, as estimated in the literature, varies from 200 to 600 ohms. Hence, Equation 7.5, which is based on the assumption of equal structure and stroke path surge impedances, must be used with caution. An Improved Equation If it is assumed that the structure and its connections have no effects upon the stroke current, which is equivalent to saying that the lightning channel impedance is large compared to the equivalent surge impedance Z as seen from the top of the structure, and, if it is further assumed that a step (rectangular) wave Iu(t) is injected by the stroke into the structure and ground wire, as shown in Figure 7.5, three equal voltages waves e(t) will travel away from the structure top along the three paths identified previously. GROUND WIRE 3e(t-T PHASE CONDUCTOR T= TRAVEL TIME = H/v (a) LIGHTNING STROKE ON A TRANSMISSION LINE STRUCTURE Figure 7.5 (b) REFLECTION LATTICE DIAGRAM FOR TRAVELLING WAVE CALCULATIONS Lightning and Travelling Waves In the initial stages, several reflections will occur at the top and base of the affected structure before the reflected waves from the adjacent structures return to the observation point, i.e., the arm of the structure. The following analysis is carried out for the time period preceding the return of the reflected waves from the adjacent structures. The voltage wave e(t) is equal to the product of the stroke current and the equivalent surge impedance: Z Z e(t) = Zlu(t) -2 s 1+21 (7-7) lu(t) As illustrated in the lattice diagram of Figure 7.5, the wave arrives at the structure arm after a time interval, h/v, where v is the wave velocity in the structure (0.7 to 0.9 the velocity of light). At the same instant, a voltage wave Me(t-h/v) proceeds along the phase conductor as a result of the retarded electric coupling with the ground wire. M is the coupling factor between the phase conductor and ground wire. The voltage wave e(t) moves down toward the base and is subject to a first reflection from the structure ground resistance. The reflected wave /3e(t) proceeds upwards and arrives at the structure arm after a time (2H-h)/v. It is then partially reflected back towards the base of the structure. This reflected wave causes a voltage a/3e(t-2H/v-h/v) at the structure arm, while the refracted waves, moving along each side of the ground wire, induce a voltage wave M(l- a)/Se(t-2H/v+h/v) on the phase conductor. The resultant voltage V(t) across the insulator strings can be calculated with the help of a lattice diagram of the type shown by Figure 7.5(a). This calculation leads to the following expression: 'l-M)u(t-T) + V(t) = ZI °o r E °J j=i L r eJ [u (t+x-jT) + [a - M(l-a)J u(t-r-jT) L (7-8) where a and /3, the reflection factors at the top and base of the structure respectively, are given by the following expressions: 1-21 9 s a = -r2----------- + (7-9(a)) 21. R s - Zs R s + Zs (7-9(b)) where r is the wave travel time from the top to the arm of the structure (r=h/v) and T is twice the wave travel time from the top to the base of the structure (T=2H/v). Surge Impedances Equation 7-8 can be expanded to include the wave exercise will not be undertaken here since it is offer additional insight into the travelling wave applicability of the surge impedance concept will to eliminate a major weakness of this approach. reflections from adjacent structures. This not essential to this study and does not technique illustrated above. Instead, the be discussed and an attempt will be made Equation 7-8 requires calculation of the reflection coefficient and coupling factors, which in turn, are derived from surge impedance values. Thus, the accuracy of Equation 7-8 in predicting the voltage stress across the insulator strings is dependent on the degree to which these factors accurately represent the fundamental parameters of the physical process and on the accuracy of the expressions used to represent these parameters. Since the travelling wave calculations have assumed constant surge impedances and lossless metallic conductors and structures (real-valued surge impedance), neither of these requirements are adequately satisfied. This subject will be expanded upon in Sections 7.4 and 7.5. It should be emphasized 7-10 that the constant surge impedance assumption is not necessarily valid in all problems, since surge impedance generally varies with time and is dependent on the transmission line structure configuration and shape of the stroke current wave. This dependency was shown by a number of authors [5,13-15,17-19] who analyzed the response of simple structure configurations (cylinder, cone) and various current waveshapes. For example, the surge impedance of a cylindrical structure to a rectangular wave current was determined [15] as: Zs = 60 ln[v^(ct/a)j (7-10(a)) where c t a isthe velocity of light, isthe time, isthe cylinder radius. If the rectangular wave is replaced by a ramp wave, the surge impedance becomes [17]: Z s 60 In /let a a 2ct + a (7-10(6)) 2ct If the transmission line structure is represented as a conical structure, then the surge impedance, when energized by an arbitrary wave current, has been shown to be constant [17]: Z s 60 1 n r' —r (7-11) Lsi n6j where 8 is the half-angle of the cone. This constant surge impedance is in reasonable agreement with experimental measurements using model structures with a conical shape [17,20]. The foregoing discussion suggests that if time-dependent surge impedances (transient or dynamic surge impedances) are used instead of the less exact constant surge impedances, accurate solutions eguivalent to those obtained from field theories can be attained using the travelling wave technigue. This is convincingly demonstrated in a discussion of the field theory approach described in [5]. The travelling wave technigue was used to develop a simple lightning back flashover computer program (STRIKE) for this EPRI project. The main feature of this program is that the stroke current wave is assumed to be made of a succession of several curvilinear segments of three basic types, suitably displaced in time. The rectangular, ramp, and exponential wave types were used on a variety of structures represented by their equivalent cylindrical representation. Typical results related to the problem illustrated in Figure 7.6, are shown in Figure 7.7. Because of the extensive work required to improve the capability and flexibility of the program in accounting for actual conditions such as variable structures, counterpoises, soil breakdown, etc., as discussed in Sections 7.4 and 7.5, no significant effort was made to enhance the program. Thus, the usefulness of the program is essentially academic. 7-11 IMPEDANCES (IN £}) EQUIVALENT CYLINDER GROUND WIRES H=30m STRUCTURE GROUNDING SYSTEM Figure 7.6 CASE NO. 7" 12 R CO 15 2 hOO oo 15 3 400 550 15 h 100 550 15 0.<l Typical Structure Top Potentials g 00 TIME (jus) Figure 7.7 Z 1 Illustration of a Typical Problem 0.2 Zt s 7.3.3 ELECTROMAGNETIC FIELD MODELS As long as the lightning performance predicted by approximate methods was satisfactory, no need was seen for a more rigorous approach. Early in 1950, however, the lightning outage rates of several transmission lines were found to be about 10 times those predicted by design. This anomaly stimulated considerable research work and as a result, an advanced electromagnetic field theory was proposed [5,13-15]. This theory validated the analytical model based on the surge impedance concept, provided that the concept was extended to include time varying and wave shape dependent surge impedances, as previously discussed. Unfortunately, it failed to explain the high lightning outage rates. Nevertheless, the field approach remains one of the most rigorous theoretical methods to describe the lightning response of transmission line structures. Two theories have been developed. The field-cancellation method, introduced by Wagner [13], and the loop-voltage method, originally employed by Lundholm et al [5] and then by Wagner and Hileman [14-15]. Both theories are based on the fundamental electromagnetic equations of Maxwell. The loop-voltage method is simpler and lends itself to practical calculations at the expense of slightly less accuracy than the field cancellation method. The loop-voltage method is explained in detail in [5] and [14] and will only be described briefly. It is of interest to note that the field cancellation method is based on the determination of a distributi on of charges which produce a quasi steady-state electrostatic field (called "cancellation field") opposite to the electromagnetic ("forcing") field produced by a set of travelling waves. The loop-voltage method, on the other hand, ignores the effects of charges because the line integral of an electric field (or voltage) about a closed loop is independent of the electric charges. Electric Field Produced by a Moving Uniform Distribution of Charges If the uniform distribution of electric charges shown in Figure 7.8 is considered and if it is assumed that these charges are moving along the cylindrical conductor at a velocity v equal to or less than the velocity of light c, the following argument can be made. Figure 7.8 Moving Wave of Electric Charges Since electromagnetic waves in free space travel with the velocity of light, the effect of an elementary charge, dq=qdx, located at point M(x,0) at the instant t would only reach the observation point P(w,y) after a time delay r/c where r is the distance between M and P. Hence, the electric field at P is retarded and the potential at point P is called the retarded potential. The preceding observation suggests that it is possible to compute the retarded potentials at any point and time if the distribution of electric charges as a function of time is known. 7-13 From electromagnetic field theory, the retarded vector potential A is related to the current I injected into a long cylindrical conductor as follows (see Figure 7.8): I(x,t-r/c) dx r A = (7-12) Where I represents the current vector flowing along the cylindrical conductor (|I|= I = qv) and r is the distance between the conductor element dx and the observation point P. It has been assumed that the current wave I is rectangular as illustrated in Figure 7.8. This assumption will be used throughout unless indicated otherwise. The vector potential at point P and time t is axial (no radial component) and is caused by the charge distribution extending from the origin o (x = 0) to the point M (x = z) reached by the current wave at the instant t-r/c. Therefore, the retardation effect can be introduced as follows: _ UoI / % Jo dx______ /(w-x) 2 + y2 (7-13) where Z = c-v U ± c TO2 + R2 + w (7-H) and U = ----- (vt-w) c+v -V (7-15 (a)) c-v c+v (7-15 (b)) The integration expression: A of s i nh Equation 7-13 is straightforward and + sinh"1 [^] results in the following (7-16) The vector potential A can be used advantageously to determine the electric potential difference between two points. As stated, the sum of the electrostatically induced voltage around a closed path is zero. Mathematically, this is equivalent to: 1 Folcv Vpath 4tt(c + v) dx (7-17) + R2 Computation of the voltage stress developed across an insulator string consists of an evaluation of Equation 7-17 along a closed integration path, including the insulator string as one portion of the path. In the following, the integration will be carried out in steps involving various portions of the closed path. This integration is necessary in the axial direction only, since the vector potential is axial and thus, has no radial component. Hence, If PI and P2 are two points located on a line parallel to the current carrying conductor (see Figure 7.8), the potential difference between these two points is determined from Equation 7-17 as: V Pi -V P2 dolv kn s i nh -1 Mi R 7-14 s i nh U RJ (7-18) where Uj and U2 are defined by Equation 7-15(a). If v is equal to the velocity of light, then Equation 7-18 should be expressed in logarithmic form. Thus for v = c: V Ct " W! P (7-19) Ct - W2. At this stage, it is of interest to note that any point outside the sphere of influence, i.e., the sphere centered at O with radius ct, is not influenced by an electromagnetic wave. Consequently, the potential along a portion of the path lying outside the sphere is zero. Thus, when PI , P2 or both are outside the sphere of influence, wj , W2 or both must be substituted by the cutoff values, as illustrated in Figure 7.8: w = ± t/c2t2 - y2 where the negative sign applies to a point P located to the left of the origin in Figure 7.8. Determination of the Currents in the Ground Wire and Structure Through repeated application of Equation 7-17 along a suitable closed path, it is possible to determine the current Ig flowing on each side of the ground wire and the current Is moving downward in the structure (see Figure 7.9). If it is assumed that the surface of the earth is an equipotential surface (zero soil resistivity) and that the structure and ground wire are perfect conductors, the value of Vpath in Equation 7-17 along the path JFK shown in Figure 7.9, must be zero since it only involves perfect metallic circuits and a segment outside the sphere of influence (segment FK). Therefore: #*‘-/$rdx'° (7-20) Noting that the currents shown in Figure 7.9 produce only tangential fields, Equation 7-20 is readily solved using Equation 7-18. For simplicity, it is assumed that the current wave velocity in the structure and ground wire is the velocity of light. Equation 7-19 can then be used instead of the more complex equation, 7-18. Integration along the JFK loop (see Figure 7.9) leads to: I Poc g 4tt 1 n ct-a ct-a + 1 n ct+a ct+a ■I s 4u 1n ct~b ct-8 = 0 (7-21) where PocAtt = 30 (a+a)2 + b2 = c2t2 (3+b)2 + a2 = c2t2 after a time t such that a«ct, Equation 7-21 reduces to: 'g [6°'"T-J-A60(7-22) The terms in brackets represent the transient surge impedances, introduced earlier (see Equation 7-10(a)). The bracketed term on the left is defined as the transient surge impedance of the ground wire, and the one on the right is the structure transient surge impedance. 7-15 Finally, the sum of the structure and ground wire currents must be equal to the total lightning stroke current 1^, i.e.: 21 g + I s = T (7-23) t Ig and Is are now determinable from Equations 7-22 and 7-23. It is interesting to note that these two equations are the same as would be derived using the surge impedance concept. GROUND WIRE STRUCTURE Figure 7.9 Integration Path of Vector Potential Voltage Across the Insulator String With the currents in the structure path must be selected to determine shown in Figure 7.10 is selected to induced in the phase conductor is insulator voltage was estimated to and ground wire determined, another closed integration the voltage across the insulator string. The loop OGBDFO perform this calculation. It is assumed that the current negligible. The effect of this induced current on the be about 8% on the conservative side [14]. GROUND WIRE GROUND WIRE PHASE CONDUCTOR (a) VIEW ALONG TRANSMISSION LINE Figure 7.10 PHASE CONDUCTOR (b)VIEW TRANSVERSE LINE Voltage Across Insulator String 7-16 TO TRANSMISSION Because the path field along GB is is zero since all voltage along the string. Therefore, V s GB is outside the sphere of influence, the line integral of the electric zero. Similarly, the sum of the line integrals along paths OG, BD and FO conducting elements are assumed to have zero resistivity. Thus, all the closed loop OGBDFO appears across the path DF, i.e., across the insulator If Vs designates the voltage stress across the insulator string, we have: -Vpath (7-24) Since path DF is orthogonal to the direction of flow of all currents, the line integral along this path is zero. Thus, only three integrals in Eguation 7-24 need to be calculated to determine the insulator voltage. This calculation is straightforward and leads to the following result: V s = 60 I In f - 30 I In g b s (7-25) ~r ct-h This eguation is valid for t > h/c (wave travel time from O to F) and for t < (2H-h)/c (time for the reflected wave to return to F). For t > (2H-h)/c, it is necessary to consider the original and the reflected wave currents. Reflected Wave Concept Thus far, the field theory has been described without introducing the concept of reflected and refracted waves. In principle, it is possible to continue the analysis without introducing this mathematical technique. What is needed is a definition of the mathematical constraints such as boundary conditions, which apply to the problem being investigated. In this analysis for example, the earth is assumed to be a perfectly conducting and therefore equipotential surface. This defines one boundary condition. Another constraint was developed when it was tacitly assumed that no charge accumulation can occur in the system. Furthermore, the stroke path is assumed to be of infinite impedance. Therefore no current originating from the ground wire or structure can flow back through this path. It is possible to develop the relations which account for all of these constraints and to solve the set of equations thus developed to determine the insulator voltage as a function of time. The reflected travelling wave technique leads to the same result however, in a very elegant manner consisting of repeated application of the approach previously outlined (for the specified first time interval) but including a suitable time lag to account for the finite velocity of electromagnetic waves. The principle of the reflected wave technique is depicted in Figure 7.11. A mirror image of the system is created and identical but opposite waves are produced by the images. This artificial arrangement satisfies the boundary conditions, since the earth's surface will always remain at zero potential; initially because no wave has reached the surface and later because both the positive (original) and negative (reflected) waves reach each point of the surface simultaneously. However, a point close to the top of the structure will not have its potential reduced by the negative or reflected wave before a time interval of about 2H/c, i.e., twice the wave travel time along the structure. This is illustrated in Figure 7.11. 7-17 TRAVELLING WAVE _____FIELD THEORY CAUSED BY ORIGINAL WAVES DOWN IN Structure (2H-h)/c time CAUSED BY FIRST REFLECTED WAVE (FROM IMAGE STRUCTURE) (a) METHOD OF IMAGES Figure 7.11 (b) TYPICAL STRUCTURE POTENTIAL Wave Reflection and Refraction Technique If the mirror image is hidden, the incident wave appears to be completely reflected when it reaches the base of the structure. When this reflected wave reaches the top of the real structure, it produces a set of new waves which will travel along the ground wire and down the structure in the same manner as the primary current wave, but delayed in time. Calculation of the magnitude of these new current waves uses a method similar to the method used to determine Ig and Is. That is, application of the constraint that the line integral of the electric field along a closed metallic path must be zero. In this case, however, there are primary and reflected current waves to consider in the equations. This process continues indefinitely because there can be no accumulation of charge in the system. Strictly speaking, the division of an upward wave at the top of the structure, into a refracted wave moving along the ground wire and a reflected wave proceeding to the base of the structure, varies with time. However, after the first reflection at the top of the structure, this division can be considered constant with good accuracy. In this case, the application of the field theory equations becomes equivalent to the use of constant reflection and refraction coefficients. Insertion of a Finite Structure Ground Resistance The field theory can readily be expanded to include the case where the structure has a finite ground resistance which can be represented by a lumped element. An effective method to account for the presence of the resistance Rs is to create a new system of current waves at the junction point between the structure base and its grounding system of magnitude - trls. cr is defined as: 2R a = -i— (7-26) s s where Zs is the structure transient surge impedance applicable at the time considered. - 7-18 Other Enhancements By application of a variety of semi-empirical formulas, it is possible to account for effects which were not considered in the preceding field theory approach. For example, corona in the ground wire can be simulated by increasing the radius of the ground wire by an appropriate amount [5]. Also, the effects of induced currents in the phase conductor can be approximated by the introduction of coupling factors [15]. Further refinements such as the inclusion of upward streamers, have also been proposed [21]. 7.4 UNCERTAINTIES DM THE ANALYTICAL MODELS OF A STRUCTURE It is legitimate to question the usefulness of efforts aimed at developing new refinements such as those described previously, particularly when the controversial nature of some of the fundamental assumptions made to construct electromagnetic field theories are carefully reviewed. Some of these assumptions are: a b c d e f - Lightning stroke modelled as an ideal current generator Lossless conducting elements Zero resistivity earth Concentrated or lumped structure ground resistance Time-invariant ground resistance Equivalent geometrical representation of a structure Although important, the first assumption will not be discussed here since it involves the formation mechanism of the lightning discharge, a very controversial subject which is well outside the scope of this work [1]. Thus, the following discussion will concentrate on the last five assumptions. 7.4.1 LOSSLESS CONDUCTING ELEMENTS All of the previously described analytical models assume that the ground wire, the phase conductors and the metallic elements of the transmission line structure are lossless conducting elements. The practical consequence of this assumption is that waves travel in the conductors at the same velocity and without shape distortion, i.e., a rectangular wave impressed at the structure top remains rectangular when it reaches its base. This property also exists in lossy conductors satisfying the "Heaviside condition" which implies that the ratio of the conductor series resistance to series inductance (per unit length values) be equal to the ratio of the leakage (or shunt) conductance to its shunt capacitance (distortionless conductors). When applied to steel conductors such as the elements of a structure, ground wires, or counterpoises, a lossless representation is difficult to accept because of the relatively high conductor series impedance, particularly during transients. Thus, a lossy line satisfying the Heaviside condition is a more appropriate assumption. Unfortunately, the conductor parameters are frequency-dependent and consequently, it is impossible to select conductor characteristics which satisfy the distortionless line condition throughout the frequency spectrum experienced during a lightning stroke. Moreover, the theory of lossy transmission lines leads to a surge impedance and propagation constant which are complex numbers, whereas real values are obtained with the lossless line approach. Finally, another significant source of losses and further computation difficulties may exist because of the nonlinear properties of steel elements which are subject to hysterisis and magnetic saturation. 7-19 The difficulty of the problem may be further compounded by the presence of nonuniform current carrying elements, i.e., elements having a position dependent surge impedance. Although it is possible to analyze lossy conductors by field theory or by the distributed circuit approach, the consideration of nonlinear effects and/or continuously nonuniform or inhomogeneous lines leads to extremely complicated and extensive computations. Such efforts, where possible, would be unproductive if the added accuracies provided by the analytical refinements were only marginal. Unfortunately, the effects of many of the preceding parameters are unknown or highly uncertain. Thus, it is only possible to speculate about the effectiveness of the contemplated enhancements. However, it is reasonably safe to state that the effects of electromagnetic losses not accounted for in the present theory are at least an order of magnitude less than the overall response and could be approximated by the inclusion of appropriate resistances at specific locations of the system. This is because the response of a transmission line structure is largely determined by the system components present within a span of the faulted structure. 7.4.2 EARTH RESISTIVITY The assumption of zero resistivity is often used in electrostatic and electromagnetic problems and allows application of the widely known "method of images". Unfortunately, it is incompatible with the previous statement concerning the localized effects of a lightning stroke on a structure, and is in direct conflict with theoretical and experimental evidence that soil in the vicinity of a buried conductor injecting current into the earth assumes a potential near the conductor potential with respect to remote ground. This situation is depicted in Figure 7.12. Typically, if the structure base dimension is 15 m (50 feet), the earth potential is still in the order of 5 to 10% of the structure potential rise at 100 m (328 feet) from the center of the base. REMOTE GROUND (0 VOLT) POTENTIAL OF BASE OF TOWER X-—- EARTH SURFACE ^ POTENTIAL EQUIPOTENTIAL SURFACE AT ABOUT 0 VOLT. DISTANCE (b) EXTENDED MIRROR IMAGE CONCEPT EARTH SURFACE POTENTIAL PROFILE Figure 7.12 Earth Resistivity and Earth Potentials 7-20 Earth surface potentials will vary with time according to a function strongly related to the current stroke. An immediate consequence of this is that it is not possible to apply the convenient "flat mirror" technique, used to create images of the original system, including travelling charges of opposite polarities. There is a surprising absence of research work or discussion on this subject, in the literature related to power engineering. This is particularly surprising in view of the extensive research work conducted on analytical models based on less significant approximations. Definitive conclusions on the effects of finite earth resistivity will have to await the development of an appropriate analytical model to account for this effect. In theory, it may be possible to use a technique similar to the method of images. However, the mirror will not be fl at and the images will be distorted and time-varying in order to keep the reference surface at zero potential at all times (see Figure 7.12). It could be more appropriate to select a surface which, during steady-state conditions, is at a specified low percentage of the structure potential and use it as the zero potential reference surface. Another solution could be to solve the problem in the frequency domain as discussed in greater detail in Section 7.5. 7.4.3 ANALYTICAL MODELS OF EXTENDED GROUNDING SYSTEMS The discussions of the preceding section are even more applicable to long grounding systems since earth surface potentials will remain at high values even at large distances from the structure. However, since this aspect of the problem has already been discussed, the following discussion will address the subject of the transient response or impedance of a long grounding system, such as a counterpoise. The steady-state impedance of a counterpoise was discussed in Chapter 5. Several useful design charts are presented in Volume 2 of this report. It is evident that the lossless transmission line approach is not valid because of the conductor shunt conductance, sometimes designated as leak ance, which plays a major role in the dissipation of current into the earth. In addition, the internal series resistance of the counterpoise can hardly be ignored. As explained, waves will travel at a fraction of the velocity of light typically 0.2 to 0.8 and will be subject to distortion and exponential decay. The transient response of counterpoises has been measured and analyzed [22,23,24]. In a more recent work [18], an attempt was made to account for the frequency dependent nature of counterpoise parameters. This work has been shown to be reasonably accurate through good agreement between calculated and measured results [25]. A similar approach, described in [26], neglects the effects of frequency on the counterpoise parameters. However, the approach allows for soil ionization effects as will be discussed here and has been treated in [27]. For comparison, the indicial transient impedances of a counterpoise as derived by Bewley [23], Devgan, and Whitehead [18] and Annenkov [26] are given: Bewley Z(t) = i-E e 8 •at k=1 (2K-1) 2rr2 cosYkt + f a L2Y. 7-21 ^k Yk " 2a 1 JI s i nY,k t 1 (7-27) where a = 1/2RC£ 1 2 I 1 (2K-1) 27T2 LC£2 1 R2C2&2 (2K-1)tt 2£/LC and ^ R C L is is is is the the the the counterpoise length (meters) dc ground (shunt) resistance (ohms) counterpoise capacitance per unit length counterpoise inductance per unit length. Devgan and Whitehead (assuming constant parameters) Z(t) = ^ [o„74e_1-1at + 0.26e"0-07at] + R [l-e"2at] (7-28) where all variables are as defined previously. In addition the following expressions were used to determine the counterpoise parameters: ,T u ohms C = iTeo(er + l)^ln - 1j farads "meter henrys/meter d = meters /2ah where a is the counterpoise radius, h its depth of burial, e0 the permittivity of free space, er the relative permittivity of soil, m soil permeability and p soil resistivity. Annenkov (response to the ramp function i(t) = It) ■2gt ■et -Jip-. 1 -e (7-29) where (8 = l/2£0erP r = H /2v (v velocity of propagation in the soil) Rc is the steady-state dc resistance of the counterpoise, including the effect of soil breakdown (ionization). Its expression is given by: Rc ■ ' 4L 2r/h 7^=1 + r J - 2 2 where -L- F 1-..l-.e"2et 1 + £EC L 28t J -8t 28t t/T 1 -e 3 and Ec is the soil critical breakdown gradient. 7-22 Although there is room for further refinements by inclusion of the effects of frequency dependent parameters and soil ionization, it is apparent that counterpoises have received adequate attention with regard to transient response. However, we are not aware of any work which uses this analytical work to examine actual lightning performance problems. 7.4.4 SOIL IONIZATION MECHANISM When subjected to an impulse current with a high crest magnitude, soil may be subject to localized electrical breakdowns at points of maximum potential gradient, i.e., near the edge of ground conductor elements. This phenomenon was discussed in Chapter 5 in the context of steady-state currents in concrete. The subject was also discussed in Chapter 10 where it was concluded that soil breakdown is rather infrequent in high voltage transmission line structure faults. Nevertheless, it is possible that for a long section of transmission line passing through a zone of high resistivity and keraunic level, this phenomenon may need to be taken into consideration if effective mitigation steps are to be found where dictated by unsatisfactory outage rates. The apparent resistivity of homogeneous isotropic soil is constant as long as the soil potential gradient around the electrodes injecting current remains below a specific value Ec, the soil critical breakdown potential gradient or soil ionization threshold gradient. As the surge current increases with time, a value Ic is reached which causes breakdown, arc discharges in the soil [27-31], and an apparent decrease in soil resistivity. The critical current Ic is generally in the kA range and its relation to Ec or Jc, the critical current density in soil, is largely dependent on the current electrode. A typical experimental resistivity-current curve is illustrated in Figure 7.13(a). This figure clearly shows the nonlinear characteristics of soil resistivity at high impulse currents. CURRENT DENSITY IN SOIL (a) HYSTERESIS IN RESISTIVITY-CURRENT CURVE Figure 7.13 (b) ANALYTICAL MODEL BASED ON GROUNDING ROD Dynamic Resistivity-Impulse Current Curve 7-23 Liew and Darveniza [27], after carefully examining various experimental test results from several types of electrodes buried in different soils [28-30], have recently proposed an analytical model to describe the time-variant (dynamic) nonlinear characteristics of some basic forms of concentrated grounding electrodes. In the model proposed, the soil is characterized by three parameters (see Figure 7.13 for illustration): Fc Critical breakdown gradient of soil (V/m),which varies between 50 and 500 kV/m, depending on soil type (zone 1). X Soil ionization coefficient (s_l). This is the coefficient which describes the rate of exponential decay of the resistivity from its nominal value at time 0 to its value at time t (zone 3). An empirical value of 0.5x106 s_l provides good correlation with test results. Soil deionization coefficient (s-1). This coefficient accounts for the soil resistivity recovery from its value at time 0 to its nominal value (zone 2). A value of about 0.22x106 s-1 was found to give the best agreement with test results. 7 As the current injected into the ground electrode increases, the current density 3 in the immediate vicinity of the electrode surface eventually exceeds the critical density Jc. Soil breakdown occurs and, progressively, the radius of the zone where soil is ionized (zone 3) increases until it reaches a maximum radius ac corresponding to the current crest. Beyond zone 3, the resistivity of the soil remains at its nominal value PQ (zone 1). However, as the current decreases from its peak value, a buffer region (zone 2) develops as the current density is now below the critical value Jc and constantly decreasing. In the deionization zone, the resistivity recovers to its nominal value. In each of the zones just described, the soil resistivity, as a function of time, is given by the following expressions: No Ionization Zone 1 (7-30) p(t) = p0 Deionization Zone 2 p(t) = pm + (po-pj [(1-J/JC)2 (l-e"Yt)J (7-3D where P m is the value of resistivity at current 3C during the decay period. Ionization Zone 3 P(t) Poe -At (7-32) The most significant result of these impulse current experiments and predictions for concentrated electrodes is that the impulse coefficient, i.e., the ratio of the impulse resistance to the low current steady-state resistance, is between 1 and some minimum value. This ratio quickly decreases toward this minimum value (about 0.2) with an increase in the impulse current peak and with increasing soil resistivity values. This clearly suggests that during a lightning stroke with a current peak sufficient to initiate the soil ionization process, a mitigating effect with respect to transmission line lightning performance is provided by the 7-24 soil breakdown. This mitigation is more effective than previously thought because of the significant reduction in the impulse coefficient afforded by concentrated electrodes in the 10 to 100 kA current range. This possibly explains the good lightning performance of sections of lines with concentrated grounding systems installed in unusually high resistivity soils. It should be noted, however, that the minimum attainable impulse coefficient increases very guickly with the number of interconnected concentrated electrodes such as ground rods. For example, the minimum coefficient is about 0.6 with a hollow sguare of 4 rods. Finally, it is important to note that the above analytical model fails to describe the performance of concentrated electrodes with localized high current densities. In such cases, breakdown will occur along discrete linear paths and the implicitly assumed diffuse growth of soil ionization, becomes comparable to a mini-lightning stroke. This situation is also likely to occur in nonuniform soils which may present one or several paths with low critical breakdown gradient strength. Despite the foregoing limitations, it does not appear that researchers or designers have taken advantage of this useful and realistic analytical model to optimize or improve the grounding systems of transmission lines. 7.5 FREQUENCY DOMAIN APPROACH A review of the analytical methods which have been developed to predict the lightning response of transmission line structures would not be complete without a discussion of the freguency domain approach which, surprisingly, has not been widely used to predict the lightning response of a structure. As evidenced in this chapter, all analytical methods discussed in the literature have been developed based on the time-domain approach. Moreover, except for a few cases, field and model measurements have been conducted with impulse or surge generators. This observation does not appear to be justified by facts, convincing evidence, or sound judgment. Rather, our assessment of the state-of-the-art supports an opposite view which will be explained. This view is also implied in [11] and is indirectly supported by many successful applications of the freguency domain method of solving transient problems. 7.5.1 FREQUENCY EFFECTS ON ELECTRICAL CONSTANTS OF MATERIALS In the preceding section, the freguency dependent nature of electric conductor parameters was mentioned and its immediate implications were briefly outlined. An obvious conseguence of this is the effect of surge wave shape on lightning performance, since any surge can also be regarded as the superposition of many harmonic components. This effect may explain, at least partially, noticeable differences in the indicial response of the individual legs of the same transmission line structure [32]. Figure 7.14(a) depicts typical curves of the dielectric and conductivity constants of a homogeneous medium. Figure 7.14(b) shows the approximate magnitude of these variations for a common substance, pure ice, at a temperature slightly below the freezing point. 7-25 102 103 DIELECTRIC CONSTANT 104 105 DIELECTRIC CONSTANT RESISTIVITY CONDUCTIVITY FREQUENCY FREQUENCY (a) Figure 7.14 TYPICAL CHARACTERISTIC _______ CURVES_________ (b) (Hz) CHARACTERISTIC CURVES OF PURE ICE (APPROXIMATE VALUES AT ABOUT 0°C) Frequency Dependence of Material Electrical Constants This frequency dependence is briefly explained by the decrease in time for the polarizing charges to separate and participate in the current conduction as the frequency is raised from a low value to the MHz range. These curves present a convincing argument in favor of the frequency domain approach, since it is excessively difficult and time-consuming to include this frequency dependence in a time simulation, while it is easily and readily adaptable to the frequency domain approach. 7.5.2 CLARITY AND PRECISION An outstanding advantage of the frequency domain approach is provided in the clarity by which the concept of impedance and surge impedance is defined, as opposed to the sometimes ambiguous definitions given to terms such as "surge" impedance, "dynamic" (surge) impedance, "transient" (surge) impedance and "impulse" (surge) impedance, when represented in the time domain. This ambiguity and the lack of wave shape definition has resulted in serious difficulties in interpreting, comparing or repeating experimental work conducted by various researchers. For theses reasons, researchers of Electricite de France [11] decided to analyze their rocket-triggered lightning test results in the frequency domain. Some of their conclusions are very perplexing indeed. For example, as discussed in Chapter 10 of this report, they conclude that the test lattice structure response to lightning strokes was essentially that of a pure inductance of about 3 to 5 mH throughout the surge frequency spectrum. This is contrary to the surge impedance concept. However, we disagree with the conclusion reached in [11] concerning the equivalent inductance representation of a transmission line structure. 7-26 This is because firstly, only one particular structure was tested, and secondly because the variation in structure parameters with frequency was not accounted for in the conclusions. It is argued here that differences in the rate by which the structure characteristics vary with frequency could have been such that the surge impedance varied linearly with frequency in the same manner an inductance normally varies with frequency. This would also explain why the measured inductance is about half the theoretical value (see Chapter 10). Nevertheless, we are in complete agreement with the view expressed in [11] concerning the use of a "harmonic" (surge) impedance which is the ratio of the measured voltage to the impressed sinusoidal current at a specific frequency. This definition, when used in conjunction with the Fast Fourier Transform technique, allows an efficient and precise means of expressing computation and measurement results conducted in the time domain to the frequency domain and vice-versa. 7.5.3 HYBRID MODELS A practical approach to the problem of representing the lightning performance of a structure can be enhanced manyfold if the many possible refinements, which can be introduced to the components of the system, lead to a final hybrid analytical model constructed from the refined and accurate submodels. In the time domain for example, the field theory could be applied to the soil ionization and the counterpoise models and the resultant hybrid model could be solved linearly in small time steps. Obviously, such a model would raise considerable difficulties. In contrast, the frequency domain method, because of its inherent linear characteristics, is particularly adaptable to hybrid analytical models. 7.5.4 CONCLUSION It is obvious, even to the nonspecialist, that the subject of analytical models for transmission line structures hit by lightning strokes, is an exceedingly complex one. This review can only briefly discuss the subject. Although original views and ideas have been expressed in this chapter, these are essentially speculative and remain to be confirmed or rejected. Only future extensive research work can help in clarifying these issues. REFERENCES 1 - R. H. Golde (Editor), "Lightning", Volume 2, Academic Press Inc., US Edition, New York 1977. 2 - Report of joint IEEE-EEI Subject Committee on EHV Line Outages, IEEE Transactions on PAS, Vol. PAS-86, 1967, p. 547. 3 - A. Cimador, R. Fieux, B. Hutzler, "Influence des Resistances des Prises de Terre sur le Foudroiement des Pylones d'une Ligne 225 kV", E.D.F. Bulletin de la Direction des Etudes et Recherches, Serie B, No. 4, 1974, pp. 29-42. 7-27 4 - M. Darveniza et Al, "Modelling for Lightning Performance Calculations", IEEE Transactions on PAS, Vol. PAS-98, No. 6, November/December 1979, pp. 1900-1907. 5 - R. Lundholm, R. B. Finn, W. S. Price, "Calculation of Transmission Line Lightning Voltages by Field Concepts", AIEE Transactions on PAS, Vol. 76, February 1958, pp. 1271-1283. 6 - A. Greenwood, "Electrical Transients in Power Systems", Wiley- Interscience, John Wiley and Sons Inc., 1971. 7 - E. F. Koncel, "Potential of a Transmission Line Tower Top When Struck by Lightning", AIEE Transactions on PAS, June 1956, pp. 457-462. - S. B. Griscom, J. W. Skoogglund, A. R. Hileman, "The Influence of the Prestrike on Transmission Line Lightning Performance", AIEE Transactions on PAS, Vo. 77, December 1958, pp. 933-941. 8 9 - J. G. Anderson, J. H. Hagenguth, "Magnetic Fields Around a Transmission Line Tower", AIEE Transactions on PAS, Vol. 77, February 1959, pp. 1644-1649. 10- F. A. Fisher, J. G. Anderson, J. H. Hangenguth, "Determination of Lightning Response of Transmission Lines by Means of Geometrical Models", AIEE Transactions on PAS, Vol. 78, February 1960, pp. 1725-1735. 11- P. Kouteynikoff, "Modele pour le Calcul de 1'Impedance d'une Prise de Terre Longue aux Basses et aux hautes Frequences", Electricite de France, Report HM/72-04670. 12- AIEE Committe Report, "A Method of Estimating Lightning Performance of Transmission Lines", AIEE Transactions on PAS, Vol. 69, pt.II, 1950, pp. 1187-1196. 13- C. F. Wagner, "A New Approach to the Calculation of the Lightning Performance of Transmission Lines", AIEE Transactions on PAS, Part III, Vol. 75, December 1956, pp. 1233-1256. 14- C. F. Wagner, A, R. Hileman, "A new Approach to Calculation of Lightning Performance of Transmission Lines. II", AIEE Transactions on PAS, Part III, Vol. 78, December 1959, pp. 996-1021. 15- C. F. Wagner, " A New Approach to the Calculation of the Lightning Performance of Transmission Lines - III; A Simplified Method: Stroke to Tower", AIEE Transactions on PAS, Part III, Vol. 79, October 1960, pp. 589-603. 16- R. Davis, J. E. M. Johnston, "The Surge Characteristics of Tower and Tower Footing Impedances", IEE Journal, 1941, Vol. 88, Part II, pp. 453-465. 17- M. A. Sargent, M. Darveniza, "Tower Surge Impedance", IEEE Transactions on PAS, Vol. PAS-88, No. 5, May 1966, pp. 680-687. 18- S. S. Devgan, E. R. Whitehead, "Analytical Models for Systems", IEEE Transactions, Part III, Vol. PAS-92, pp. 1763-1770. Distributed Grounding 19- J. A. Selvaggi, R. J. Nigbor, R. L. Kloecker, "The Green’s Function of a Cylindrical Tower", IEEE Transaction paper No. C72-558-5, IEEE PES Summer Meeting, July 1972, San Francisco, California. 20- P. C. Buchan, W. A Chisholm, "Surge Impedance of HV and EHV Transmission Towers", Final Report No. 78-71, Canadian Electrical Association. 7-28 21- C. A Liew, M. Darveniza, "A Sensitivity Analysis of Lightning Performance Calculations for Tranmssion Lines", IEEE Transactions on PAS, Vol. PAS-90, 1971, pp. 1443-1451. 22- E. D. Sunde, "Surge Characteristics of a Buried Bare Wire", AIEE Transactions, Vo. 59, 1940, pp. 987-991. 23- L. V. Bewley, "The Counterpoise", General Electric Review, Vol. 37, No. 2, February 1934, pp. 73.81. 24- L. V. Bewley, "Theory and Tests of the Counterpoise, "Electrical Engineering, (AIEE Transactions), Vol. 53, August 1934, pp. 1163-1172. 25- M. Kawai, "Studies of Tower Footing Resistance on Transmission Lines", IEEE paper 31 CP 65-704, IEEE Summer Meeting, Detroit, Michigan, 1965. 26- V. Z. Annenkov, "Calculating the Impulse Impedance of Long Earthings in Poor Conducting Ground", Electrichestvo, No. 11, 1974, pp. 62-79. 27- C. A. Liew, M. Darveniza, "Dynamic Model of Impulse Characteristics of Concentrated Earths", Proceedings IEE, Vol. 121, No. 2, February 74, pp. 123-135. 28- P. L. Bellashi, "Impulse and 60 Cycle Characteristics of Driven Grounds", AIEE Transactions, Vol. 60, March 1941, pp. 123-128. 29- P. L. Bellashi, R. E. Armington, A. E. Snowden, "Impulse Characteristics of Driven Grounds", Vol. 61, 1942, pp. 349-363. 30- K. Berger, "The Behavior of Earth Connections Under High Intensity Impulse Currents", CIGRE Publication No. 215. 31- E. Y. Ryabkova, V. M. Mishkin, "Impulse Characteristics of Earthings for Transmission Line Towers", Electrichestvo, No. 7, 67-70, 1976. 32- E. J. Rogers, "Impedance Characteristics of Large Tower Footings to a 100 ns Wide Square Wave of Current", IEEE Paper 80 SM 662-7, IEEE, PES Summer Meeting, Minneapolis, Minnesota, July 1980. 7-29 8.1 GENERAL Only a small proportion of the electric schock incidents reported each year involve high voltage power eguipment or installations. Most electrocutions are caused by lightning or involve low voltage circuitry and household electrical devices. A recent survey conducted among North American utilities indicates that there have been no fatal accidents due to excessive touch or step voltages in the vicinity of transmission line structures [1]. A summary of the results of the survey is given in Chapter 11. This by no means suggests that incidents involving high voltage installations have not occured. We are aware of several electrocutions involving grazing cattle in the neighbourhood of high voltage lines and one incident with a human fatality which occured in Europe and is described in [2]. The victim, a camper, apparently was crawling out of his tent at the time a phase-to-ground fault occured at a nearby transmission line structure. It is apparent however, that only a very small number of fatalities can be attributed to high earth surface voltages near faulted transmission line structures. A recent study in Finland [29] estimates the probability of an accident to be in the order of 10-8 per year per structure for 120 to 420 kV structures. The measurements and theoretical analyses conducted during this EPRI project and elsewhere [3,4] suggest that this extremely low accident rate is not a result of transmission line structure grounding design but rather is a consequence of a very low probability of human exposure to hazardous situations. More precisely, for an electrocution incident to occur, the following events must occur simultaneously: a- A ground fault at a transmission line structure sufficient in magnitude to generate hazardous step and touch potentials. b- A person must be in a dangerous position near this specific structure at the time of the fault. Because of the infrequent presence of people near high-voltage transmission line structures and because a fault at any specific structure is very unlikely, the electrocution incident is an extremely low probability event. This favorable situation may not exist at some exposed locations along transmission lines or may change in the future because of increased joint use of transmission line right of ways in densely populated urban centers and progressively higher fault current magnitudes and probabilities of fault occurence. It is the project engineer's responsibility to determine the transmission line locations which present high human exposure risks and, if necessary, to design a suitable grounding system to control hazardous earth surface voltages. While this chapter is intended to help the engineer in assessing the nature and severity of a ground potential rise hazard, it also describes a methodology for designing a suitable grounding system for earth surface potential control based on the methods developed in Chapters 4, 5 and 6. 8- 1 8.2 THE ELECTROCUTION EQUATION It is now well established that the duration and magnitude of the current traversing the human body determines the severity of an injury. The well known Dalziel electrocution equation, I5 = p/v^, where p is a constant proportional to body weight, t is the time duration of the electric shock and Ij-, the rms value of the maximum permissible body current, is the reference equation used by the majority of North American engineers to determine the safety of power system grounding installations [5,6,7]. The electrocution equation, as applicable to humans, can only be determined empirically by extrapolating experimental data obtained from tests on four-legged animals and primates. The electrocution equation is also designated as the fibrillation threshold of current as a function of duration. Dalziel's equation [5,13,14] is based on the statistical analysis of published experimental data and, since the minimal data which exist on electrocutions is subject to interpretation, it is not surprising to note that there is major disagreement among researchers concerning the exact nature of the electrocution equation and that different fibrillating threshold levels are in use in other countries [7]. It appears from more recent research work [9,10,11,12] that Dalziel's equation may not be sufficiently conservative in the 0.3 to 3 second fault duration range and may be overly conservative for body weights in excess of 60 kg (133 lbs). The electrocution equation is of prime importance in the design of safe grounding and protective relaying systems. This equation describes the relationship between the fibrillating threshold body current, shock duration, and body weight during exposures to voltage sources of various magnitudes and frequencies. Although a literature search shows an urgent need for further research work to establish more suitable electrocution equations, it is not within the scope of this EPRI project to develop, recommend or criticize new or existing equations. However, the various theories which have been developed concerning these equations are summarized in the next section. Thus, it is the role of the design engineer to select the most suitable equation, based on the applicable standards, guides [8], existing rules and regulations. The information in this section gives a brief summary of several research works aimed at determining the most suitable form of the relation between the maximum permissible body current I^ (also designated as the minimum fibrillating current) and the various influencing parameters. These parameters are not all well-known and it is clear that further work is required to fully understand the significant parameters. However, three variables, body path characteristics, current frequency, and current duration have been shown by experimentation to be closely interrelated to the safe body current magnitude. Therefore, the general form of the electrocution equation can be written as: Ib = <i>(p,f >t) (8-1) where Ib is the maximum permissible body current in amperes rms (for a given body path p) which will not cause death to more than x % of a sample population having similar body path characteristics. f is the current frequency in Hz. t is the duration of the fault in seconds. p is one or several parameters characterizing the body path followed by the current. 8-2 Dalziel's equation was established for power frequency current (50 to 60 Hz) durations between 8 ms and 5 s and a percentage x of 0.5 % of the population. The body path is characterized by the constant p which is proportional to body weight. 8.2.1 THE ELECTROCUTION MECHANISM The passage of electric current through living biological specimens can produce stimulation of excitable tissue at low current intensities, or coagulation of protein and burning and charring at high current intensities. Although in many cases, persons have survived high currents despite severe injuries, most electrocution fatalities occur during low-intensity current shocks because of asphyxia (respiratory arrest) or heart fibrillation (cardiac arrest). Fibrillation, more common than asphyxia as a cause of death, is the random contraction and relaxation of the heart muscle fibers in the ventricles resulting in a loss of the pumping action of the heart chambers and cardiac arrest. Unless circulation is restored (for example, by de fibrillation) no later than 3 to 4 minutes after the electric shock, irreversible brain damage occurs. Respiratory arrest occurs during the passage of current because of the contraction of the chest muscles. Generally, spontaneous resumption of respiration occurs after the shock current has ceased. For shock durations of a few seconds or less, respiratory arrest is unlikely. It is only when the arrest of respiration persists after the electric shock that a serious risk of death exists, if the victim is not resuscitated by artificial respiration. Lee [15] describes several electrocution incidents which suggest that this sustained respiratory arrest occurs only when the current path is through the respiratory center located in the lower part of the brain. Since most power frequency electrocution incidents do not involve the head, fibrillation is usually used as a reference in the design of safe power installations. Respiratory arrest is probably of interest in the case of a direct lightning strike on standing persons, a subject not within the scope of this project. 8.2.2 THRESHOLDS OF CURRENT A current in the order of 5 mA can be sensed by the tongue while 0.5 to 1 mA is the minimum current which can be perceived by the fingers. As the current increases, a tingling sensation occurs followed quickly by a painful muscular contraction. At a current of 6 to 7 mA for women and 9 to 10 mA for men (at 50-60 Hz) a threshold is reached beyond which it is no longer possible to release the energized conductor. This current is called the let-go current. For still higher currents (20-50 mA) breathing difficulties, suffocation and even asphyxia may occur if the electric shock duration lasts for minutes. Currents in the order of 100 to 200 mA may cause death by heart fibrillation if the electric shock duration is in the order of 0.5 second. 8.2.3 PARAMETERS INFLUENCING HEART FIBRILLATION The fundamental variables which govern the electrocution Equation 8-1; the magnitude, frequency and duration of the current, and its path through the human body; depend on a number of other biological parameters which, unfortunately, are statistical rather than constant in nature. Since there is only limited experimental data, it is understandable that the exact values of some of these parameters are still uncertain. The role played by these parameters is becoming somewhat clearer through research in the medical and engineering fields. The significant parameters described in this section have a direct effect on the magnitude of the current which traverses the body of a human or animal 8-3 subject to electric shocks. Other parameters have more subtle, unknown or sometimes controversial effects. In order to fully understand the interaction between these variables, it is important to describe the "strength -du rati on curve" of excitable living tissue. This is well summarized in [16]. Experimental measurements show that there is an approximately hyperbolic relation between the stimulus strength s (current density of the electric stimulus) and the duration t of this stimulus: s = f + r (8-2) Equation 8-2, which is a very well established relationship in physiology, was originally proposed by Lapicque [17]. Figure 8.1 is a plot of Equation 8-2 and shows that the threshold current density required for stimulation increases as the duration of the stimulus is decreased. There is a minimum current r (designated as the rheobase) for an infinitely long duration stimulus, below which no tissue excitation is possible. The constant a, depends on the type of excitable tissue. STIMULUS STIMULUS DURATION (s) Figure 8.1 Strength-Duration Stimulus Curve There are other equivalent forms of the strength-duration equation. Referring to Figure 8.2, the following develops an equation based on the equivalent electrical circuit of a cell. The cell membrane excitable tissue can be represented as a capacitor C shunted by a resistance R. The transmembrane potential difference is slightly less than 100 mV at rest. Following an electric stimulus which reduces this potential by about one third or more, a regenerative process takes place which depolarizes the cell and reverses the cell polarity slightly before restoring the transmembrane resting potential. When the membrane is of a muscle cell (e.g., heart muscle), contraction follows depolarization. The product RC is the membrane time constant r. A current stimulus s applied to this parallel RC circuit is related to the time constant r and to the rheobase r as follows: s r (8-3) 8-h s s STIMULUS STIMULUS (a) (b) EXCITABLE TISSUE Figure 8.2 EQUIVALENT CIRCUIT Equivalent Circuit of a Typical Cell Membrane Pearce and his colleagues [18] have shown that there is good agreement between the measured strength-duration curve and the relation predicted by Equation 8-3. Therefore, Equation 8-3 can be used instead of Equation 8-2 to predict the response of excitable tissue to electric stimulus. However, it is important to note that this response is based on the application of a single electric pulse (stimulus). When repetitive stimuli are applied to the tissue, only a sub-multiple of the pulses will stimulate the cells because of the refractory period of the cell. The refractory period is the time interval following a stimulus where any further stimuli applied to the cell will not be detected by the cell. The above considerations explain why low frequency currents (50-60 Hz) produce ventricular fibrillation more readily than higher frequency ac currents. If, at high frequencies, sinusoidal alternating current peaks are considered as a succession of short duration pulses, only a fraction of these pulses are effective in stimulating the tissue. Thus, as the frequency is increased, the magnitude of the fibrillating threshold current is increased. This phenomenon is well known by surgeons and is the basis for recently developed diathermy and electrosurgery techniques. Ventricular fibrillation can be produced by repetitive electric stimuli more readily than with a single stimulus. Experiments show that at power frequencies, the magnitude of the current required to initiate ventricular fibrillation is a minimum. A single stimulus (e.g., lightning surge) can precipitate ventricular fibrillation only when it occurs during the time interval of the ventricle pumping cycle (slightly less than 1/3 of the heart cycle) and when the stimulus is approximately twenty times higher than the threshold of excitation when the heart is at the end of the period of relaxation and filling of the ventricles (diastole). This dangerous time interval is called the vulnerable period of the heart cycle and could explain the observation that only about 1/3 of persons struck by lightning die. In diastole, an electric stimulus of the heart causes a premature heart beat if the duration of the stimulus is less than the duration of the premature beat. For a longer duration stimulus or repetitive stimuli, several premature heart beats are initiated and the beats follow each other with increasing frequency. After each premature beat, the fibrillation threshold (initially 20 times the excitation threshold) is progressively lowered until it approaches the normal threshold of excitation of the ventricle during the nonvulnerable period of the heart (i.e., diastole). 8-5 The above discussions are in agreement with experimental results [19] which indicate that there is a discontinuity of the 50 % ventricular fibrillation probability curves for electric shock durations in the order of one third the heart cycle as shown in Figure 8.3. THRESHOLD OF FIBRILLATION FOR 50% FIBRILLATING PROBABILITY NON FIBRILLATING THRESHOLD ■=- OF HEART 3 CYCLE v TIME (ms) Figure 8.3 Threshold of Ventricular Fibrillation for a 50 Kg Man The transition zone between short and long duration levels is of particular interest to power system engineers because it corresponds to time durations in the order of the normal response time of protective relays (0.2-2 s). In this zone, the relation is linear with a slope very close to -1. This suggests that for the time range 0.2-2 s, the equation It=constant (Osypka [12]) is more appropriate than the IvC=constant proposed by Dalziel [13]. In practice, however, there is little difference in the results of the two equations for durations corresponding to this transition zone. In spite of the considerable progress since the beginning of this century, the data related to electrocution are still very experimental in nature and cannot be accurately correlated with actual electrocution cases. The major difficulty is in that most meaningful shock data have been obtained under experimental conditions where the current stimulus was directed toward the target, (i.e., the heart) generally through direct stimulations of the target. In actual electrocution accidents, contact is generally made through the extremities (hands or feet) and only an unknown fraction of the total current reaches the vulnerable tissues of the body. Unfortunately, it appears that there has been no work directed at a determination of the current paths through a human body and there is no indication that efforts have been made to extrapolate any test data available from quadrupeds to humans. The scarcity of experimental data relating human and animal response to electric shocks is certainly one of the reasons for the existence of several forms of the electrocution equation as developed by various researchers working with essentially the same experimental data. 8-6 8.2.4 THE ELECTROCUTION THEORIES In the design of an electrical grounding system, one of the important aspects to be considered is the safety of the installation during fault conditions when current enters the grounding system and causes differences in potential between various accessible points of any grounded metallic structure and the soil surface in the vicinity of the grounding system. In order to assess the safety of an installation, the engineer must know the highest tolerable body current (or voltage) which must be compared with the calculated or measured currents under fault conditions. This important reference value is the current fibrillation threshold or tolerable body current usually specified by national and international standards or guides and is determined from the electrocution equation discussed previously. In Norh America, the equation proposed by Dalziel [14] is recommended by the widely used IEEE Guide No. 80 [6,7]. In other continents and countries, different equations are in use [8]. However, all equations lead to practically equivalent results and relate the magnitude of the body current to its duration without distinguishing between the points of application of the electric stimulus which is implicitly assumed to be sinusoidal and at power frequency (50 or 60 Hz). Despite some lack of flexibility of the equations in coping with the wide variations in the values of the variables encountered in practice, they are of vital importance in the design of suitable mitigation steps for effective electrical shock prevention. There are essentially four major time-dependent electrocution equations. These have been developed by Dalziel [14], Osypka [12], Geddes [10] and Biegelmeier [19]. The equations were derived empirically from a statistical analysis of the fibrillation current versus time data. A recent paper by Bridges [20] provides a good summary of these four major concepts and their related equations: Dalziel's Equation Dalziel concluded after examining results of experiments conducted by others, that the fibrillating threshold of current is inversely proportional to the square root ot the electric shock duration: Ib = p//t (8-4) Osypka's Equation Using essentially the same data as used by Dalziel, Osypka developed an equation where the fibrillating threshold is inversely proportional to the electric shock duration (except for durations exceeding 1 second where the current remains constant): Ib = d/t (8-5) Geddes' Equation Based on physiological considerations pertaining to the irritable tissue concept discussed earlier in this section, Geddes proposed the following equation where the current is also inversely proportional to the electric shock duration: Ib = a/(t-b) + c (8-6) Biegelmeier's Equation Bielgelmeier's work is also based on the irritable tissue concept. The equation proposed by him is illustrated in Figure 8.3. Bielgelmeier considers that the fibrillating threshold current is discontinuous with time: ^ = A < 0.2 rb = B > 2 < t < 1 b = C/t ; 0.2 s s (8-7) 2 8-7 s In the above electrocution equations a, b, c, d, A, B, C and p are statistical constants which depend on the degree of safety to be achieved. Since absolute safety is economically prohibitive and unpractical, the maximum safe body current is selected such that x % of the population, consisting of "standard" persons of a defined body weight, will not be subject to fibrillation (for example, IEEE Guide 80 [6] proposes 99.5 % of 50 Kg persons). Figure 8.4 shows the safe body current as a function of its duration according to Dalziel, Osypka and Bielgelmeier. BIELGELMEIER OSYPKA DALZIEL i i i i TIME (mS) Figure 8.4 Safe Body Current Versus Time As previously stated, it is not within the scope ot this EPRI project to propose an electrocution equation. Consequently, the general form of the equation as given by (8-1), will be used. 8.2.5 THE ELECTRIC SHOCK CIRCUIT Electric shocks are possible only when current flows through the body of a human or animal. This necessarily means the existence of a closed path, the electric shock circuit, which includes at least a portion of the body of the victim. 8-8 Because of the numerous activities performed by human beings and the variety of electrical apparatus used, the points of contact in an electric shock can be anywhere on the body. However, in the case of potential transmission line grounding incidents, the most common points of contact will be at the hand or foot extremities. Generally, two types of contacts are considered: step contact and touch contact. These are illustrated in Figure 8.5. This figure also shows the equivalent circuit which applies in each case. The various parts of the circuit have been identified and can be grouped in three categories: 1- Resistance due to man-made insulating materials (gloves, shoes, or soil surface covering insulating materials). 2- Body surface resistance (skin resistance). 3- Resistance of internal body tissues (heart, hands, legs, etc.). EQUIVALENT CIRCUIT ACTUAL CIRCUIT (b) STEP CONTACT TYPE Figure 8.5 The Electric Shock Circuit It has been shown previously that the fibrillation threshold depends on current passing through the heart tissues. More precisely, the current density in the heart tissues appears to be the primary factor in determining whether or not fibrillation occurs. The equivalent circuits of Figure 8.5 show that the distribution of body current between the heart (unit 5) and other parts of the body, depends on the path followed by the current and on the ratio of the heart resistance (R5) to the neighbouring tissue resistances (Rg for touch contact, R4 and Rg for step contact). The magnitude of the total body current is dependent on the voltage across the contact areas and on the various resistances of the electric shock circuit. 8-9 No literature has been found concerning research on the distribution of shock currents through the human body or the resistance of current paths through the human body. Consequently, the detailed equivalent circuits of Figure 8.5 are of academic interest only. In practice, the body path is represented by one lumped equivalent resistance (see Figure 8.6) and total body current, not heart current, is used to define tolerable current levels. For example, in Dalziel's equation, only body weight is a factor. The safe body current is independent of the current path. As can be concluded intuitively from Figure 8.5, this is not the case. Moreover, according to Louchs [24], a man can withstand a foot-to-foot current about 25 times higher than a hand-to-foot current. It is important to remember, however, that a nonfatal but painful foot-to-foot current might result in a fall with possibly serious consequences. Thus, more sophistication can not be justified without further research and great care must be exercised in the selection of the biological constants. Simple, conservative equations appear to provide the most reasonable approach to shock hazard determination. Body Resistance It is useful to qualitatively consider the effects of the various body resistances of Figure 8.5. Firstly, it is clear that the resistance of body current paths such as legs or hands, is inversely proportional to the cross-section of the paths. This leads to the conclusion that body resistances of a child are higher than those of an adult. This is an important fact observed experimentally by Whitaker [21] and means that for the same voltage across the contact areas, less current will flow through the body of a child. It should be noted that for identical situations, the areas of contact for a child are usually less than for an adult (hands or feet). Consequently, the contact resistances are also higher. Despite the lack of conclusive data relating to children, it may be argued that although the fibrillating threshold current for a child is less than for an adult, the same voltage limits might apply to both the adult and the child, because of the inherently higher-child body resistance. The preceding considerations raise an important but still unresolved question. What is the relationship between body weight and fibrillation threshold? It is known that the minimum fibrillation current increases with body weight. According to Dalziel [5], this current is proportional to weight. In contrast, Geddes has concluded that the fibrillation current varies with the square root of the body weight. However, based on the elec trophy si ological concept described earlier, one can conclude that for bodies of the same general proportions, the response of excitable tissue is identical provided that the current density in the tissue is identical. The current density criterion [20] appears more appropriate for extrapolation purposes and leads to a fibrillating current proportional to the 2/3 power of body weight. The 2/3 power relationship is derived from the observation that for a given shape of different weights, cross-sectional area varies as the 2/3 power of the weight. Since most experiments were made on quadrupeds, extrapolation of the results to bipeds based on the body weight criterion will introduce unknown errors. The body current density factor also raises difficulties due to the unknown distribution of current within the body of the victim. Again, uncertainties prevent us from selecting the most appropriate assumption. Therefore, the more simple relations, although somewhat inaccurate, remain a reasonable choice. Figure 8.6 shows a simplified equivalent circuit of the electric shock path. In this circuit, the whole body path is represented by a single resistance. Several experiments [20-23] show that this resistance is a function of the voltage applied across the contacts and skin condition at the contact areas. Excluding skin contact resistance, the resistance of the internal body tissues and organs with current flowing between hands and feet is in the order of 300 to 500 ohms. At relatively low voltages (< 250 V) the contact resistance of dry intact skin is in the order of 100 kilohms/cm2. However, if the skin is wet or not intact, the contact resistance may be only 8-10 a fraction of a kilohm and will become considerably less effective in limiting body current. Since skin resistance, dominant under normal conditions, can easily be reduced to a value comparable to the body internal resistance, various researchers have suggested lower bound values of combined skin and internal body resistance. Typical suggested "worst case" values range between 500 and 1500 ohms. gloves r ^body tower shoes covering (a) TOUCH VOLTAGE Figure 8.6 (b) STEP VOLTAGE Simplified Equivalent Electric Shock Circuit It should be noted that touch or step voltages in excess of 100 volts are not uncommon in the vicinity of faulted transmission line structures. At these higher voltages the body resistance exhibits a negative, nonlinear characteristic as shown in Figure 8.7 from Freiberger [25]. Body resistance has also been shown to decrease with time if the duration of contact is in the order of a few seconds or more and the voltage is above 50 volts [22,26]. This is a result of skin breakdown, often evidenced by the blisters, which is caused by the heat generated from the current flow. Since transmission line fault durations are generally less than 0.5 s, the IEEE Guide 80 recommended body resistance (i.e., 1000 ohms) is believed to be reasonably conservative in this context of transmission line grounding systems. Resistance of Gloves and Shoes Relatively new dry shoes are probably the best protection against electric shock hazards. It is probable that many theoretically dangerous situations for bare-footed persons go unnoticed by persons wearing good insulating shoes. According to an extensive experimental survey conducted by Electricite de France [27], the resistance of dry new shoes varies between 50 and 200 megohms. When wet, the resistance drops dramatically to the kilohm range, with values as low as 800 ohms. If the shoes are not new and are wet, the resistance drops below 500 ohms in many cases. Thus, it may be prudent to ignore the effect of shoe resistance, especially when safety precautions are considered at locations such as recreational or camping areas where bare-footed persons, especially children, are very likely. 8-11 Since the use of gloves is very uncommon except when safety measures are clearly necessary, the glove resistance shown in Figure 8.6 should generally be assumed to be zero. AVERAGE WITH FINGERS (10cm2 CONTACT) ___ AVERAGE WITH FULL HAND (100cm2 CONTACT) ------ AVERAGE VALUES •yr ZONE OF MEASURED ^ VALUES VOLTAGE STRESS (volts, rms) Figure 8.7 Body Resistance Versus Applied Voltage (Redrawn from [25]) Surface Soil Resistance of Feet For the purpose of circuit analysis, the human foot is usually represented by a conducting metallic disc or rectangular plate. The contact resistance of the foot with the soil is generally neglected. In some cases, a 5 to 15 cm (2 to 6") layer of crushed rock (or similar semi-insulating material) is spread on the surface above the ground grid to provide a series resistance between the natural soil and the feet of a person standing above the grid. This practice is beneficial only when this surface layer resistivity is significantly higher than the natural top soil resistivity (say 5 times or more). Otherwise, a significant grounding system current will flow in this surface layer and the earth potentials will be significantly different from the potentials computed assuming a uniform soil. Assuming that the surface layer resistivity is Ps and the thickness of this layer is e, this soil-to-foot series resistance Rf becomes: Rf = ps f (8-8) 8-12 where S represents the area of the metallic disc or plate representing the foot. With a 6 cm thick layer and a 0.02 m2 foot contact surface, Rf is 3 ps, a value commonly used in substation grounding. Ground Resistance of a Foot If a person is standing well away from a grounded structure so that no mutual resistive coupling exists between the grounding electrodes represented by the feet and the structure ground system, the resistance to remote ground of the feet must be taken into consideration in the electric shock path equivalent circuit. A typical situation where this factor must be considered occurs in the case of transferred potentials, as illustrated in Figure 8.8. In this case, the series resistance of the foot, not only includes the surface layer (when it exists), but also the natural surrounding soil. This resistance is then more appropriately designated as the foot resistance to remote ground. Volume 2 of this report provides charts to determine the foot resistance when the soil consists of a surface layer and an equivalent infinite deeper stratum. TO SOURCE Figure 8.8 Ground Resistance of Feet The Safety Criterion It is now possible to define the condition which will warrant that the body current is not a fribrillating threshold current for at least x % of the population. If Vf is the calculated touch voltage (hand-to-feet), the body current i must be less than the fibrillating threshold 1^: Vt ' = Rb + iR‘- !b = If Vs is the calculated step voltage (foot-to-foot) we similarly have: V i=---------------< i = 4>(p,f,t) (8-10) Rb + 2Rf 8-13 For example, if Dalziel's equation is used, assuming a body resistance of 1000 ohms and no surface gravel, Equation 8-9 becomes: Vt 1000 p /t where p = 0.116 for 99.5 % of a population consisting of 50 kg adults. 8.2.6 SAFE IMPULSE OR SURGE CURRENTS Essentially, everything said so far regarding safe current levels applies equally well to very short (microseconds) or long duration currents (seconds) with few exceptions. That is, the safety criteria of Equations 8-9 and 8-10 are valid regardless of the shape and duration of the electrocution voltage. Unfortunately, the values of the biological constants which affect the electrocution equation are not well known when impulse, lightning or power system surge currents are involved. Several direct and indirect lightning strike and power system surge electrocution incidents indicate that a person can survive very short duration currents (1 to 1000 ms) of 20 to 100 A, as illustrated in Figure 8.9 which is based on capacitor discharge incidents reported by Dalziel [28]. This indication is confirmed in Figure 8.10 which shows the effect of frequency on perception current, also reported by Dalziel [5]. With increasing frequency, the perception of current changes from tingling to the sensation of heat. DANGEROUS ZONE SAFE ZONE TIME CONSTANT (microseconds) Figure 8.9 Nonfatal Surge Capacitor Discharge Currents (Redrawn from [28]) 8-14 ___ PERCENTILE 0.5 ___ PERCENTILE 50 ------ PERCENTILE 99.5 FREQUENCY (Hz) Figure 8.10 Perception Current as a Function of Frequency (Redrawn from [5]) From a limited number of electric shock incidents involving the discharge of high voltage capacitors, Dalziel [28] concluded that the energy criterion, l2t=constant, is applicable for impulse or surge currents. An uncertainty exists concerning the appropriate value of body impedance when subject to high magnitude surge currents. Not only is the body impedance reduced because of the high voltage level, it is also shunted by a capacitance. The value of this capacitance is not well known although suggested values range from 100 to 500 pF. This capacitance appears to play a mitigating role by diverting a major portion of the current which would otherwise flow through the body tissues. There is clearly an urgent need for extensive research concerning electric shocks caused by surge or impulse currents. However, until better information is available it seems prudent to adopt conservative low body resistance values, i.e., about 300 ohms or less. It is important to note that the transmission line structure grounding system provides much better protection against electric shocks from lightning than a tree or any other poorly grounded structure. Thus, the engineer need not be concerned about shock hazards from the lightning incident itself but rather, should investigate the possible power system fl as hover resulting from a lightning strike on the transmission line structure. If engineering analysis indicates that mitigation measures must be applied, these should be based only on the power frequency current effects. 8-15 8.3 PROBABILISTIC CONSIDERATIONS Safety criteria, although derived from extensive experimental data, must recognize that absolute safety remains an unachievable ideal and that some degree of risk is present in many aspects of life. The concept of "relative safety" is already embedded in the electrocution equation, 8-1, which prescribes the fibrillating threshold current for a certain percentage of the population. In other words, it is recognized that there is always a very small percentage of the population which, for various reasons; extreme fatigue, heart disease, etc., will be vulnerable to currents harmless to the great majority of the population. Although it is the function of engineers to indicate the practicality and economical cost of achieving the different levels of safety, it is the role of governemental and standard committees to establish the maximum acceptable level of risk. Consequently, the probabilistic methodology presented in this section should not be used without prior knowledge of its applicability, and conformance to the regulations, standards or company directives which are in force. 8.3.1 THE PROBABILITY OF AN ELECTROCUTION INCIDENT The occurence of an electric shock in a power system network, particularly at a transmission line structure, requires the simultaneous occurence of two unfavourable and, generally, unrelated events: a- Occurence of a power system ground fault resulting in a circulation of all or a fraction of the fault current into earth through a transmission line structure. b- Presence of a human being in an exposed position near the structure. It is important to note that such an occurence may be hazardous to life only if the fault current magnitude and duration are such that the current generated in the body exceeds the permissible value. In other words, the incident may be fatal only if the stress function (body current) exceeds the strength function (threshold of fibrillating current). Since high voltage transmission line faults are relatively infrequent and generally of very short duration (<0.5 s), the two events, a and b, are rarely coincident. This is a significant factor in the apparent absence of fatalities from electrocution due to transmission line structure faults. 8.3.2 THE STATISTICAL PROBABILISTIC APPROACH In Finland, where the average soil resistivity varies between 1000 and 10,000 ohm-meters [29], most attempts to meet the grounding requirements of the Finnish electrical safety code (1957) have been condemned to failure because of the difficulties in obtaining low power system ground resistance at realistic costs. This unacceptable situation prompted an extensive study which resulted in a new safety code [30] based on a systematic probabilistic analysis. This technique permits the design of a grounding electrode with a predicted yearly maximum number of electroction incidents. This approach could not have been implemented without extensive statistical data related to power system outages and population exposure to transmission line structures in Finland. 8-16 The large and increasing number of transmission line structures and the very low observed electrocution accident probability associated with these structures would appear to justify a similar probabilistic approach in North America. Because this approach appears to be applicable to transmission line grounding system design, the methodology developed in [30] and presented in [29] will be summarized in this section. However, it must be kept in mind that the probabilistic concept presumes that: a- Statistical data are available in order to determine the probability of the various events leading to the accident. b- A maximum acceptable number of yearly electrocution accidents at any transmission line is defined. There is sufficient data on power system outages to permit a realistic evaluation of the probabilities associated with transmission line faults. However, as yet, there is no North American guide which attempts to define an acceptable probability of an electric shock accident at a transmission line structure. Determination of Accident Probability The probability that the fibrillation current threshold will be exceeded when the human body conducts an arbitrary current i, is given as: R = F(i) f(i)di (8-11) where F(i) represents the body strength function, i.e., the cumulative distribution of maximum permissible body current for a given population sample. f(i) represents the body stress function, i.e., the distribution of the possible body currents at a fault current site. The body strength function can be determined statistically from physiological tests on animals. By analyzing and extrapolating these results to humans, it is possible to derive heart fibrillation threshold values as a function of time for various probability percentages (0.5 %, 1 % etc.). By assuming a Gaussian distribution and taking into account the noncritical heart-time interval, Karkainnen and Palva [29] determined by an iterative method the current and voltages corresponding to the different probabilities of heart fibrillation. Their results show that the constant energy law (e.g., Dalziel's equation for a 0.5 % probability) agrees well with their function in the time range of 0.2 to 5 seconds. For times less than 0.2 s or greater than 5 s, the function is essentially constant. This function can be written as: [oTtT ♦ °'25]4.(i) ; for t < 0.56 s ; for t > 0.56 s F(i) (8-12) <Mi) where t is the time in seconds and <t>(\) represents the cumulative distribution function of a (0,1) normalized Gaussian distribution. This formula assumes that the Gaussian distribution is valid for the whole current range. 8-17 The body strength function F(i) varies with the duration of the electric shock but is well defined for each specific current duration. However, the body stress function f(i) varies widely from one grounding installation to another and is dependent on the grounding system arrangement, soil structure, power system configuration and characteristics at that location. Thus, F(i) must be determined at each site where a grounding system is to be designed. This task may be cumbersome since it requires the calculation of body current at various locations (and in various positions) throughout the grounding system. Furthermore, it is necessary to express the magnitude of these various calculated currents as a probability function. In practice, this is feasible only through computer simulation. Alternatively, it is possible to adopt a conservative approach which assumes that regardless of location, a person will be subject to the maximum computed body current i0. In this case the function f(i) reduces to the Dirac (or impulse) function 5(i0) of the current i0, and the probability R of Equation 8-11 becomes equal to the body strength function F(i0). Once F(i) and f(i) have been determined, the probability R that the fibrillating current will be exceeded if a person is exposed to a shock voltage at the site under investigation is obtained using Equation 8-11. However, the final probability P of an accident depends on the probability H of a power system fault involving the grounding system at the site during the observation period T (usually one year). This probability depends primarily on the average frequency of ground faults X involving the site during the observation period T. It is also necessary to consider the fact that a person is not always present in a possibly hazardous area. This is accomplished by including in the analysis the probability W of a person being located in a hazardous position during the period T. The final probability of an accident is therefore: P = R H W (8-13) which expresses, in probabilistic form, the previously defined fundamental conditions required to cause a potential hazard to life. In practice, H is determined from the observed statistical distribution of ground faults and W is derived from population density statistics and the average exposure of an individual to the site under consideration. Based on their statistical observations and using the fundamental concepts of the probabilistic theory [31], Karkainnen and Palva [29] showed that the product H.W can be expressed as: (8-14) H W = Z P. (l-qnk) k=1 K where n _ _-AT Pk “ e _ _ , (XT)k kl------- a+b/n a2 + (b/n)2 q - i----- tT— +-------*—— 2T2 8-18 and a is the duration of the fault, b is the time spent by the person at the site. n is the number of equal time intervals into which b has been divided. X is the average frequency of ground faults at the site during the observation period T. In the above, q represents the probability that the hazardous voltage will not be observed during one period of exposure, while is the probability of having k ground faults during the period T. Equations 8-11, 8-13 and 8-14 can be used to accurately estimate the final probability of an accident. It is apparent however, that computations of this probability will require the determination of several difficult statistical distribution functions. It is also evident that if a decision is taken to proceed with the probabilistic approach, careful examination of the detailed approach (briefly described here) and statistical data as presented in [29,30] is necessary. However, it is genuinely difficult to assess the effectiveness and/or appropriateness of a probabilistic approach to safety, based on sophisticated computations, particularly when our limited knowledge of the physiological effects of electric currents is considered. Further research work in this area is certainly justified. For the present, depending on the nature and complexity of the problem being considered, it may be concluded that simplified computations, based on conservative assumptions, are best used to determine the three fundamental probabilties of Equation 8-13. The simplest analysis is obtained when values of these probabilities are known from previous work or assumed equal to known conservative values. For example, the probability W of having someone at the site in an exposed position may be taken as 1. An interesting alternative could also be provided through the use of empirical relations to determine R, H and W. These relations would necessarily be based on extensive statistical data derived from observations at actual installations of the type being considered; in this case, high voltage transmission lines. 8.3.3 THE SELECTIVE DETERMINISTIC APPROACH The major disadvantage of the probabilistic approach is that it can be used only if certain statistical information related to the site under consideration is available. Moreover, even if the necessary information is gathered, an acceptable probability limit is necessary to achieve safe design objectives. Finally, since the statistical data related to faults on transmission line structures do not usually make any distinction between individual structures, the probabilistic approach may lead to situation where protective measures are taken at structures remote from populated centers, while other structures located in densely populated areas, may not be equipped with mitigation measures because of significantly lower fault probabilities. For the above reasons, it may be argued that it is preferable to use a selective deterministic approach rather than a probabilistic approach, especially when safety around transmission line structures is being investigated. The terms "deterministic" and "selective" are used to emphasize the fact that electric shock hazard is considered as certain at only some particular structures where the risk of exposure is estimated to be particularly high. 8-19 Structures located in densely populated centers or in areas such as schoolyards, parks or recreational areas are obvious examples of vulnerable transmission line structures. This approach has the advantage of permitting the analysis of any structure on an individual basis. Consequently, it allows decisions to be based on the merits of each case. Another advantage of the approach is the standardization which may be developed in identifying the types of transmission line structures which must be thorougly investigated from a safety point of view. For example, three classes of structures may be considered: a- Structures located in areas where the density of the population is less than x per square kilometer (low exposure). b- Structures located in areas where this density is greater than or equal to y per square kilometer (y>x) (moderate exposure). c- Structures located in specific areas where the risk of exposure of humans to the structure is unusually high (schoolyards, recreational parks, etc.). An intermediate class may be added, recognizing that in moderate exposure areas, safety investigation is only required at those structures which may rise to a potential above a specific value. For example this new class is created by stating that: d- Structures which are erected in areas where the density of the population is between x and y per square kilometer (x<y) and which, during ground faults, may be raised to a potential exceeding U kilovolts. Many variations to the above illustrative rules could be proposed and could include some aspects of the probabilistic approach described earlier. The final decision regarding the optimum choice remains the responsibility of the design engineer who, based on the present and future standards, will have to decide whether or not improvement and/or mitigation measures are required for the grounding system of the structure being considered. 8.4 SAFE DESIGNS AND MITIGATION TECHNIQUES As shown in Chapter 5, the earth potential in the vicinity of the grounding system of a transmission line structure is directly proportional to earth resistivity p and to the current I discharged to the soil by the grounding system. Hence, the touch or step voltages which exist in the vicinity of a transmission line structure are also proportional to the resistivity P and current I, since they are a function of the difference in earth potential between two specific points. Consequently if U is used to designate a stress voltage (touch or step), then the following relation is applicable: U = A p I (8-15) Based on the same uniform soil conditions, the resistance R of the structure is also proportional to P and can be written as: R = B p (8-16) In the above two relations, A and B are constants which generally depend on the configuration of the grounding system. A and B can be readily determined from the equations presented in Chapter 5. Equations 8-15 and 8-16 will now be used to illustrate the characteristics of the equation which is used to evaluate the level of safety at a transmission line structure. 8-20 8.4.1 EVALUATION OF THE SAFETY LEVEL Equations 8-9 and 8-10 show that the body current i can be expressed in the general form: i = (8-17) Rb + where U is the stress voltage (step, touch, etc.), Rj-, is the body resistance between the points of contact and Rp is the total resistance of all protective, semi-insulating materials in the current path outside the body. As previously stated, this resistance may be zero in the worst case or may include a relatively high skin resistance, shoe or glove resistance or resistance of a thin layer of gravel on top of the natural low resistivity top soil. In the following, the resistance R^ is assumed equal to 1000 ohms and Rp is selected as 1.5PS where ps is the resistivity of a thin layer of gravel. As discussed in Section 8.1, this value of Rp corresponds to the parallel resistance of both feet of a person standing on the gravel. Therefore, using Equation 8-15, Equation 8-17 can be rewritten as: A p I 1000 + 1.5PS (8-18) The magnitude of the current I is dependent on many variables, as explained in Chapter 6. However, it can be expressed to a good approximation as follows [32] (see also Figure 8.11): V |R(1+Zs/Zg) I (8-19) where V is the magnitude of the phase to ground voltage. Zs is the equivalent current source impedance as seen from the phase conductor at the faulted structure (Zs=rs+jxs). -g is the equivalent ground wire (and counterpoises, if any) return path impedance as seen from the faulted structure (Zg = rg+jxg). For the sake of simplicity in the following discussion, the impedances Zs and Zg are assumed to be pure inductances, i.e., Zs ~ jxs and Zg ~ jxg. According to Equation 8-16 and 8-19, Equation 8-18 becomes: ; = ______________ A p v ........................ ......... (1000 + 1.5PsK/b2 p2(l + Xs/Xg)2+ (8-20) xs2 The above body current is safe only if it does not exceed the value of Ij-, in Equation 8-1. If Dalziel's Equation 8-4 (with p=0.116) is used to determine I^, this condition becomes: _________ A p V_________ < Vx 2 + (1 + x /x )2B2p2 s s g' 116 + 0. 174Ps Vt 8-21 (8-21) Although this inequality has been derived from certain specific conditions, it is of tremendous importance since it incorporates the key paramaeters used to mitigate ground potential rise hazards in the vicinity of the structure. Z Figure 8.11 s Approximate Equivalent Transmission Line Circuit Before proceeding with the next section, the reader is reminded that it is implicitly assumed that Ps is several times larger than the top soil resistivity p such that the response of the grounding system is essentially unaffected by this relatively thin, high resistivity surface layer. Typically, Ps is assumed to be 5 times P or greater since, for lower values, its effectiveness decreases drastically and its usefulness becomes questionable. 8.4.2 IMPROVING SAFETY AROUND TRANSMISSION LINE STRUCTURES A close examination V and the equivalent as design constants. Modern transmission primary relays, and of the variables which intervene in (8-21) reveals that the line voltage source impedance xs can not be easily modified and should be considered Also the fault clearing time is, for all practical purpose, a constant. lines are usually designed to clear a fault in less than 0.1 s with the 0.5 s with the back-up protection. However, careful attention should be paid to the possibility that the fault remains uncleared, should the steady state fault current be less than the relay setting. This situation will generally occur at structures not equipped with a ground wire or continuous counterpoise and with high ground resistances. With long clearing times t, the right-hand side term of (8-21) reduces substantially and the safety criterion will often not be satisfied. In this case, the first corrective measure is to install sufficient ground conductors at the structure to ensure its resistance drops to a sufficiently low value to cause a fast relay response. 8-22 At this stage, we are left with the following design parameters to control the safety around the structure: p Average resistivity of local volume of soil Ps Resistivity of thin layer of covering gravel Xg Equivalent neutral current return path impedance A, B Configuration of the grounding system The A and B functions are dependent on the extent and geometry of the grounding system. It is important to note that normally the extent is more or less represented by B and the geometry is essentially represented by A. Because these parameters can be modified at the structure to varying degrees, they the four fundamental techniques which can be economically used to control step potentials at the structure. Depending on the circumstances, some techniques effective than others. Often, however, a combination of two or more techniques are to achieve optimum results. designate and touch are more necessary In this section, techniques related to the first three parameters are discussed. The technique concerning the extent and geometry of the grounding system is described in the next section, because it usually represents the final mitigation measure when the other techniques fail to achieve the required level of safety. In order to better assess the effectiveness of the various hazard mitigation techniques, it is essential to note that according to (8-21), the complexity of the mitigation task increases with the line voltage V except when the extent of the grounding system increases (B decreases) and the source impedance (Zs = jxs) decreases at the same or a higher rate. This situation usually occurs when four-legged lattice towers are used. With tubular or single-shaft structures, supplemental grounding conductors are required to provide a similar effect. In addition, it should be noted that when the source impedance can be neglected in (8-21), the level of safety (left-hand side of inequality) becomes independent of the average soil resistivity. This remark is of fundamental importance as will be explained. Finally, when xs is not negligible, the left-hand side of (8-21), which is proportional to body current, is a function of the average soil resistivity and exhibits a maximum for some unfavourable soil resistivity, Pu. This corresponds to the positive root of the first derivative of this function. In practice, this means that there is a particular resistivity value which will cause a maximum hazard for a given installation. Several of these curves have been presented in [33,34]. Figure 8.12 shows typical curves. Chemical Treatment of Soils Depending on the value of the source impedance Zs the body current curve will approximate one of the typical shapes shown in Figure 8.12(a). Examination of these curves reveals that chemical treatment of soils to decrease earth resistivity, is effective in reducing body currents only if the final resistivity value achieved is to the left of the theoretical maximum of the curve. In fact, as is illustrated in this figure, a resistivity decrease could increase the body current. These observations underline the importance of exercizing caution in requesting chemical treatment of the soil around a transmission line structure. In summary, it can be said that the effectiveness of the treatment increases with the source impedance, as seen from the structure. In practice, it is of little benefit to apply chemical treatment at a structure with a ground resistance several orders of magnitudes larger than the source impedance, as seen from the structure. 8-23 Except in simple cases, a final decision should be taken only after careful examination of the computation results. It must be remembered that any chemical treatment remains effective only for a certain time period and must be reapplied as required. For this reason it is not recommended, except in exceptional cases. (b) TYPICAL DESIGN RESULTS (a) TYPICAL CURVE SHAPES Figure 8.12 Body Current Versus Resistivity Effects of a Thin Layer of Gravel It is possible to improve the safety around a transmission line structure by adding a thin layer of high resistivity gravel on top of the natural soil. This method is appropriate when the resistivity of the gravel is several times larger than the top soil resistivity. Otherwise, it may influence the response of the grounding system to a degree that it is no longer accurate to compute the grounding system response without including the layer of gravel in the soil model. The use of gravel as a means of decreasing the hazard around a structure suffers two other drawbacks, serious enough perhaps to limit its use to a small number of locations where preventive maintenance is available. The first problem associated with the use of gravel results from the wide variations in the resistivity of gravel, depending on the origin of the material. According to the measurements described in [34], the resistivity of gravel varies not only with moisture content but is also dependent on the location from which it was originally extracted, as shown in Figure 8.13. This problem is alleviated if the type of gravel is clearly indicated in the design specifications. The second difficulty associated with the use of gravel is the likelihood of deterioration of the original insulating properties through a progressive accumulation of soil and wind blown debris in the voids between the gravel material. This should be of no concern if regular maintenance is provided at the protected site or if other low maintenance materials, such as asphalt, are used. Installation of Ground Wires or Counterpoises Chapters 3, 6 and 10, recent publications [32,33], and a number of charts in Volume 2 of this report clearly show that overhead ground wires or counterpoises can dissipate a significant portion of the total fault at a structure. 8-24 Consequently, ground wires and counterpoises represent a very effective means of reducing the electrical shock hazard at critical structures. The effect of installing such neutral conductors is a reduction in the impedance Zg 0f the return current paths, as can be concluded from (8-21). If the transmission line is not equipped with overhead ground wires or if it has insulated ground wires, a counterpoise connecting the structure to several adjacent structures offers an effective hazard reduction technique, since it works as a stress equalizer among the connected structures. In fact, safety is generally improved at all structures compared to the case where each structure is isolated from the others. The only disadvantage is that any ground fault at one structure will involve all other structures in the group so connected. The equations of Chapter 6, computer program PATHS, and the charts of Volume 2 provide several methods which can be used to accurately compute the effectiveness of a proposed design. Furthermore, Chapter 3 provides a good illustrative example of the significant effects of the overhead ground wires. RESISTIVITY OF GROUND WATER # 57 CLEAN CRUSHED GRANITE (GARNER, N.C.) WASHED LIMESTONE (GULLIVER, MICHIGAN) WASHED LIMESTONE iPRESQUE ISLES, MICHIGAN) CRUSHER RUN (GRANITE (GARNER, N.C.) RESISTIVITY OF WATER (ohms-meter) Figure 8.13 Resistivity of Gravel (Redrawn from [34]) 8.4.3 GROUND POTENTIAL MITIGATING CONDUCTORS If the usual grounding system of a transmission line and its overhead ground wires (if they exist) are not sufficient to provide a safe environment at critically located transmission line structures, the installation of some special additional ground conductors should be considered to reduce step and touch voltages. 8-25 A practical and widely used arrangement [3,35] of such conductors consists of a ring or a square loop buried at a shallow depth around the structure base. If necessary, several concentric grading rings buried at progressively increasing depths can be installed in difficult cases. These rings are usually designated as ground potential control rings or loops. Some typical ring arrangements are shown in Figure 8.14. PROFILE PROFILE RING No. 2 (100 ■2 RINGS 1 3 RINGS RING RING RING LATERAL DISTANCE (a) FOUR-LEG LATTICE TOWER Figure 8.14 %) LATERAL DISTANCE (b) SINGLE-SHAFT TOWER Typical Ground Potential Mitigating Rings The ring provides beneficial effects on the overall performance of the grounding system as follows: a- It reduces the ground resistance and, in most cases, the potential rise of the structure. b- It reduces the steep potential gradients which exist near the buried part of the structure. c- It reduces the earth surface potentials in the vicinity of the structures. These effects result in touch and step voltage reductions as illustrated in Figure 8.14. Examples of the mitigating effects are presented in Chapters 3 and 10 and in numerous charts in Volume 2. Additional examples are shown in [3,35]. The example of Appendix B of [35] displays the results of field measurements made on a single ground rod equipped with a variable number of concentric grading rings, placed progressively deeper in the ground. Field measurement and computation results related to actual installations are described and discussed in [3] and in Chapter 10 of this report. The main conclusion drawn is that touch and step potentials can be substantially reduced by a single ground potential mitigating ring, optimally dimensioned. Further reductions are obtained by the installation of additional rings at progressively increasing distance and depth. Theoretically, the touch and step potentials can be reduced to an inconsequential value by the use of a sufficient number of rings. However, it is important to note that the effectiveness of each additional ring decreases very quickly after the installation of the first ring. 8-26 8.4.4 OTHER SAFETY MEASURES Other means of improving safety in the vicinity of a critically located transmission structure include the installation of an insulating fence and/or covering the soil surface around the structrure with a thick layer of a good semi-insulating material such as asphalt. This layer must be properly inclined as to avoid water accumulation. These techniques may not always be practical or aesthetic in park or recreational areas. The presence of a fence is not necessarily an obstacle to everybody and may attract children. Thus, whenever possible, techniques such as the installation of ground wires, counterpoises or ground potential rings should be used because they provide invisible and effective mitigation. REFERENCES 1 - EPRI Project RP 1494-1 Interim Report, "Transmission Line Grounding - Survey of Utility Practices" by Safe Engineering Services Ltd., Montreal. 2 - A. Goubet, "Sur les Risques pour les Personnes Resultant de la Montee en Potentiel du Sol au Voisinage Immediat d'un Support de Ligne Electrique en Gas d'Ecoulement a la Terre d'un Courant de Defaut", Internal Report ST/AL/MC, Direction du Gaz de 1'Electricite et du Charbon, Paris, December 6, 1974. 3 - E. A. Cherney, K. G. Ringler, N. Kolcio and G. K. Bell, "Step and Touch Potentials at Faulted Transmission Towers", IEEE Transactions on PAS, Vol. PAS-100, No. 7, July 1981, pp. 3312-3321. 4 - F. Dawalibi, W. G. Finley, "Transmission Line Tower Grounding Performance in Nonuniform Soil", IEEE Transactions on PAS. Vol. PAS-99, No. 2, March/April 1980, pp. 471-479. 5 - C. F. Dalziel, "Electric Shock Hazard", IEEE Spectrum, February 1972, pp. 41-30. 6 - IEEE Guide 80, "Guide for Safety in Substations Grounding", Institute of Electrical and Electronics Engineers, New York, 1976. 7 - F. Dawalibi, M. Bouchard, D. Mukhedkar, "Survey on Power System Grounding Design Practices", IEEE Transactions on PAS, Vol. PAS-99, No. 4, July/August 1980, pp. 1396-1405. 8 - V. Yablonski, B. Shatkovska-Karpa, "Allowable Touch Voltages in the COMECON Standard Proposal", Electrichestvo, No. 7, 1980. 9 - J. E. Bridges, K. Rogler, J. Wingfield, R. Kaleckas, "Electric Shock Prevention Investigation", Final Report HTRI Project E6373, Contract No. CPCS-76-0093, Oct. 1977. 10- L. A. Geddes, L. E. Becker, "Principles of Applied Biomedical Instrumentation", 2nd Edition, John Wiley and Sons, 1975. 8-27 11- G. Biegelmeier, K. Rotter, "Electrische Widerstrade Und Strome in Menschliken Korper", E Und M 99, 1971, pp. 104-109. 12- P. Osypka, "Quantitative Investigation of Current Strength, Duration and Routing in AC Electrocuti on Accidents Involving Human Beings and Animals", Technical College of Braunschweig, Brunswick, W. Germany, 1966/SLA Translation Center, TT-6611470. 13- C. F. Dalziel, "Threshold 60-Cycle Fibrillating Currents", AIEE Transactions, Vol. 79, Part III, 1960, pp. 667-673. 14- C. F. Dalziel, W. R. Lee,"Reevaluation of Lethal Electric Currents", IEEE Transactions, Vol. IGA-4, No. 5, September/October 1968, pp. 467-476. 15- R. H. Golde, "Lightning", Academic Press, New York, 1977, Chapter 16 of Vol. 2, by W. R, Lee, pp. 521-543. 16- G. L. Ford, L. A. Geddes, "Transient Ground Potential Rise in Gas Insulated Substations; Assessment of Shock Hazard", IEEE Paper 82 WM 010-7, IEEE PES Winter Meeting, New York, January/February 1982. 17- L. Lapicque, "Definition Experimentale de 1'Excitabilite", Proc. Soc. Biol. 1909, 77:280:283. 18- J. A. Pearce et Al, "Stimulation With Ultrashort Duration Pulses", PACE 1981. 19- G. Bielgelmeier, W. R. Lee, "New Considerations on the Threshold of Ventricular Fibrillation for AC Shocks at 50-60 Hz", IEE Proc. Vol. 127, No. 2, Pt. A, March 1980. 20- J. E. Bridges, "An Investigation of Low-Impedance Low-Voltage Shocks", IEEE Transactions on PAS, Vol. PAS-100, No. 4, April 1981, pp. 1529-1537. 21- H. B. Whitaker, "Electric Shock as it Pertains to the Electric Fence", Underwriters Laboratories, Bulletin of Research No. 14, pp. 3-56, 1939. 22- M. S. Hammam, R. S. Baishiki, "A Range of Body Impedance Values for Low Voltage, Low Source Impedance Systems at 60 Hz", Paper 82 SM 476-0, IEEE PES Summer Meeting, San Francisco, July 1982. 23- A. Sances, Jr., et Al, "Effects of Contacts in High Voltage Injuries", IEEE paper 81 WM 217-9, IEEE PES Winter Meeting, Atlanta, February 1981. 24- W. W. Loucks, "A New Approach to Substation Grounding", Electrical News and Engineering, May 15, 1954. 25- J. Freiberger, "Der Electrische Widerstand Des Men Menschliche Korpers", Berlin, Springer, 1934. 26- W. B. Kouwenhoven, "Effects of Electricity on the Human Body", Electrical Engineering, Vol. 68, pp. 199-203, March 1949. 27- "Recherche de 1'Influence des Chaussures sur les Dangers Presentes par les Tensions de Pas", Certificat No. 162763, Electricite de France. 28- C. F. Dalziel, "A Study of the Hazards of Impulse Currents", AIEE Transactions on PAS, October 1953, pp. 1032-1043. 8-28 29- S. Karkainnen, V. Palva, "Application of Probability Calculations to the Study of the Earthing Voltage Requirements for Electrical Safety Codes", Reprint from SAKHO, Electricity in Finland, No. 11, 1974. 30- Publication Al-74 of the Electrical Inspectorate, "The Electrical Safety Code", in Finnish or Swedish, Helsinki, 1974. 31- A. Papoulis, "Probability, Random Variables and Stochastic Processes", London, 1965. 32- F. Dawalibi, D. Mukhedkar, "Ground Fault Current Distribution in Power Systems - The Necessary Link", IEEE Paper A 77 754-5, IEEE PES Summer Meeting, Mexico City, July 1977. 33- F. Dawalibi, D. Bensted, D. Mukhedkar, "Soil Effects on Ground Fault Currents" IEEE Transactions on PAS, Vol. PAS-100, No. 7, July 1981, pp. 3442-3450. 34- G. E. Smith, "Resistivity of Some Surface Coverings Used in Substation Yards as a Function of Resistivity of the Wetting Agent and as a Function of Time After Wetting", Internal Report, April 1978, Ref. KJ 1W18, Carolina Power and Light, N.C. 35- The Electricity Authority of New South Wales, "Earthing Handbook", Australia, June 1975. 8-29 FIELD MEASUREMENT TECHNIQUES The design of optimum ground electrodes is critically dependent on soil resistivity at the site. Thus, it is essential that resistivity be accurately determined to allow an accurate prediction of grounding system performance. It is usually desirable to check the design calculations by measuring the ground resistance of the electrode after construction. For transmission lines, standard grounding systems are normally used all along the line except at a limited number of critical locations which require special investigations. Consequently, measurements at most sites are limited to verification that the resistance values are within design limits and to determine if additional grounding is required. Field resistance and resistivity measurements are therefore an integral part of the engineering work related to transmission line grounding. 9.1 GROUND RESISTANCE MEASUREMENT The principle of ground electrode resistance measurements is illustrated in Figure 9.1. This figure shows, in schematic form, an earth electrode to be tested, a return electrode, which can be any other ground electrode sufficiently distant from the test electrode, and a power source which is used to pass current between the two electrodes. The measurement objective is to determine the rise in potential of the test electrode with respect to remote ground as a result of the test current I. The principal difficulty encountered in ground resistance measurements is in locating a suitable return electrode. To facilitate proper measurements, the electrode under test and return electrode must be completely "isolated" as far as any metallic conduction paths or mutual earth resistances are concerned. If an existing structure is to be used as a return electrode, assurances must be obtained that there are no metallic connections, however indirect, between the two grounds. In practical terms, this assurance is usually very difficult to obtain. Consequently, a new return electrode is normally installed for the purpose of the tests. The requirements for the return electrode are dictated, to a large extent, by the characteristics of the grounding system being measured, and, to a lesser degree, by the soil conditions. In general, the measurement problems become more difficult as the ground impedance of the grounding system decreases, as the dimensions of the grounding system increase, and as the level of ambient electrical noise and soil resistivity both increase. The problems described are generally of minor concern in carrying out resistance measurements on transmission line structure grounds. This is because of the small physical dimensions, generally remote location, and relatively high resistance of structure grounds as compared to substation or large building grounds. The connection of a continuous overhead ground wire creates an additional set of problems that will be discussed later in this chapter. 9-1 If it were possible to readily locate an ideal remote ground reference point for any ground electrode resistance measurement, and if the isolation of the test and return electrodes could always be assured, resistance measurement would be a very simple process as illustrated in Figure 9.1. One voltage measurement between the grounding system and remote ground for a given test current would yield the desired result. In practice however, neither an ideal remote reference point or an assurance of isolation is obtainable and a measurement technique must be used which compensates for, or identifies deviations from the ideal arrangement so meaningful results can be obtained. The fall-of-potential method is the most widely used measurement technique in this regard and is the principal method discussed in this chapter. I STRUCTURE BEING MEASUREDX REMOTE" RETURN CURRENT ELECTRODE C2 "REMOTE" POTENTIAL /PROBE JL-P, Figure 9.1 Principle of Ground Resistance Measurement 9.2 TEST METHODS There are three principal methods used to measure the resistance of ground electrodes; the two-point method, the three-point method, and the fall-of-potential method [1], The fall-of-potential method is by far the most widely used in the electrical industry according to a survey conducted in 1976 [2]. 9.2.1 TWO-POINT METHOD In this method the total loop resistance of the unknown and return ground electrodes is measured. The resistance of the return electrode is assumed to be negligible in comparison with the resistance of the unknown ground and the measured value in ohms is assumed to be the resistance of the unknown ground. The technique is obviously limited to the measurement of relatively high resistance grounds, such as a single ground rod, or small ground array, provided that a low ground resistance system is readily available at the test site. 9.2.2 THREE-POINT METHOD The three-point method involves the use of two auxiliary grounds (2 and 3) of resistances designated T2 and rj. The unknown ground resistance r^ is determined by measuring the resistance between each pair of grounds and solving for rj from the equation: r - r 12 i + r 23 (9-1) 13 2 9 -2 There are several limitations to this technique which restrict its use to the measurement of small, high resistance grounds in areas where the three grounds can be located well out of the zone of influence of each other. Since neither this method nor the two-point method have any built-in checks to assure that significant errors are not being introduced into the measurements, the fall-of-potential technique has become the universally accepted test method. The two and three-point methods should only be used to obtain rough estimates of ground resistance or to perform repeated measurements for experimental purposes once the adequacy of the test configuration has been independently verified. 9.2.3 FALL-OF-POTENTIAL METHOD The fall-of-potential method of resistance determination involves the use of an auxiliary return current electrode and a series of surface potential measurements taken at increasing distances from the grounding system under test. Provided that the auxiliary electrode has been located sufficiently remote from the grounding system and coupling effects between voltage and current leads have not affected readings, the potential measurements will reach (or become asymptotic) to a level which represents the rise in potential of the tested electrode due to the test current. The general arrangement of electrodes is shown in Figure 9.2. Current electrode Cl and the fixed potential electrode PI are located on the grounding system. The remote auxiliary current electrode is designated C2. The potential electrode P2 is sequentially located at regular intervals moving away from Cl and voltage and current readings taken at each P2 location. An apparent resistance value is determined from each set of voltage and current readings. If the apparent resistance is plotted against distance, a levelling of the apparent resistance values will be observed as the potential probe P2 becomes remote from the grounding system. In carrying out fall-of-potential measurements, the voltage lead may be extended in the same direction as the current lead or at an angle, usually 900. Although less convenient, keeping the voltage and current leads at right angles to each other minimizes the mutual impedance between the leads; a very desirable situation when low impedance grounds are being measured. For transmission line structure measurements or measurements involving small substations, the unknown ground resistance is usually large enough and the zone of influence around the tested electrode small enough that mutual impedance errors are insignificant regardless of current and voltage lead orientation. When measuring the resistance of transmission line structure grounds, it is highly desirable to locate the remote current electrode in an orthogonal direction to the transmission line right of way in order to minimize the induced voltages in the test leads and equipment. This has safety as well as measurement implications because of the large unbalances which can occur in the phase conductors during ground faults. If the unknown resistance is approximately one ohm or larger and the return electrode has been adequately located, a potential profile taken from the tested electrode to the return current electrode will look similar to the curve of Figure 9.2. (see also Figures 9.5 and 9.6). A distinct flat portion of the curve, indicating a zone under the influence of neither the tested nor return electrodes, can readily be observed. The apparent resistance observed in this portion of the curve is the resistance of the tested electrode. In many cases, however, the curve obtained during fall-of-potential measurements does not resemble the curve of Figure 9.2 and interpretation procedures must be applied to make use of the data or to correct the test set-up. 9-3 FOUR TERMINAL RESISTANCE TEST SET C, V o Pi O POTENTIAL PROBE IN OPPOSITE DlRECTION STRUCTURE BEING TESTED '.i —0i i t LJ POTENTIAL PROBE tnr TRUE RESISTANCE VALUE FALL OF POTENTIAL ^ CURVES ^ X Figure 9.2 RETURN CURRENT ELECTRODE C2 0 P*o POTENTIAL PROBE DISTANCE Fall-of-Potential Test Schematic Diagram and Typical Test Results 9.2.4 THEORY OF THE FALL-OF-POTENTIAL METHOD Figure 9.3 shows an electrode E located at some distance from a second current carrying electrode G. If no current is conducted through electrode E and if E is located at a vary large distance from electrode G, or any other current carrying electrode, the self potential, Vee, (ground potential rise) of E is zero. If a current I the electrode potential rise the electrode enters electrode E, the potential rise of E, Veej is given as IRe where Re is impedance. If I = 1 ampere, then Pe = Vee = lRe = Re. Pe designates the of electrode E when 1 A flows in the electrode. Pe is numerically equal to impedance in ohms. Assume now that at some finite distance from E, a current I is passed into the soil through electrode G (E does not conduct any current). Because of the local earth potential, electrode E, initially at zero potential rise, will be at potential Veg as a result of the resistive coupling between E and G. If I = 1 A, then Veg = Peg, which is numerically equal to the so called "mutual resistance" between E and G. If electrode E also carries 1 A of current, the potential rise of electrode E will be Pe + Peg. The theoretical expressions from which Pe and Veg can be determined are complex and will not be presented here except for simple earth and electrode configurations. The more general expressions were determined in Chapter 5 for the grounding systems of transmission line structures. Derivation of the Fundamental Equations The problem is illustrated in Figure 9.3. The potential probe location x0 which leads to an apparent ground impedance (V/I) equal to the grounding system impedance Re, is required. 9-4 The current i flowing in current I injected in the return current electrode presented in this section, U e the potential probe F is neglected compared to I. At a time t, grounding system E, is assumed positive and I, collected by the G, is assumed negative. Based on the definitions and symbols the following relations hold: = Pfe 1 - P fg (9-2) = P I - P£ I e fg (9-3) Uf and Ue are the potentials, with respect to remote ground, of electrode F and E respectively. The voltage V measured by the fall-of-potential method is: V = u - u£ e f therefore I (P V - P e eg pr + fe fg (9-4) ) Pe is the potential caused by electrode E on itself when carrying a current of 1 ampere. This is by definition the impedance Re of electrode E. Therefore, Equation 9-4 can be written as: V I R e P + eg (9-5) " Pfe5 Pfg and Pfe are functions of the spacing between the electrode (E, G and F), the configurations and the soil characteristics. electrode a- Uniform Soil Let us define the following functions a, shown in Figure 9.3, (it is assumed that and x): 4> and a, <f> ^ with respect to the coordinate system and ^ are only functions of distances d = ct(d) (9-6) Pfg = <j>(d “ x) (9-7) P eg Pfe = (9-8) ^(x) According to Equation 9-5 the measured impedance R = V/I will be equal to the true impedance Re if: F> fg - P eg - P£fe = 0 that is: (j) (d - x) - a(d) - ip (x) = 0 Identical Electrodes and Large Spacinqs. If electrodes E and G are identical d is large enough such that Peg = a(d) ~ 0, then Equation 9-9 becomes: cj>(d - x) - ifj(x) = 0 (9“9) 4> = \(f and if (9-10) Hemispherical Electrodes. If electrodes E and G are hemispheres, their radii are small compared to x and d, and soil is uniform, the potential functions a, <f> and ^ are inversely proportional to the distance from the center of the hemisphere. If the origin of the axis is at the center of hemisphere E, Equation 9-9 becomes: 1/(d - x) - 1/d - 1/x = 0 (9-11) The positive root of Equation 9-11 is the exact potential probe location x0: x = 0.618 d o This is the usual 61.8% rule [9]. If the potential probe F is at location F2 (E side, see Figure 9.3) then d-x should be replaced by d+x in Equation 9-11. In this case the equation has complex roots only. If F is at location F^ (G side, see Figure 9.3) then d-x should be replaced by x-d in Equation 9-11. The positive root in this case is: x o = 1.618 d b- General Case If the soil is not uniform and/or electrode E and G have complex configurations then, the functions a, 4> and ^ are not easy to calculate. In such cases, computer solutions are generally required [3]. TRANSMISSION LINE STRUCTURE BEING \ MEASURE[K X-------- POTENTIAL PROBE RETURN CURRENT ELECTRODE Figure 9.3 Required Potential Probe Location in Fall of Potential Measurements 9.2.5 INTERPRETATION OF FALL-OF-POTENTIAL DATA In order to carry out accurate resistance measurements using the fall-of-potential method, it is very important that the theoretical limitations of the technique be understood. It is also important that the engineer or technologist conducting the measurements appreciate the effects that various deficiencies in a test set up will have on measurement data so that the deficiencies can be corrected on site and useful measurement data obtained. 9-6 As stated previously, the difficulties encountered with fall-of-potential measurements increase as the size of the ground electrode being measured increases and the resistance to be measured decreases [3]. Although it is unlikely that serious difficulties of this nature will be encountered in many transmission line ground measurements (assuming that the tests are carried out before the overhead ground wire is installed), the reasons for these difficulties will be discussed in detail. Firstly, as the resistance of an unknown electrode decreases, the voltages to be measured in the vicinity of the electrode become lower while the electrical interference usually becomes greater. This is because of increased power distribution at the large sites, be they HV transmission lines, substations or industrial facilities. For example, a large substation in an area of low resistivity could have a resistance of 0.25 ohms or less. The typical test current from a portable Megger type instrument is in the order of 40 ma. This combination results in a total induced voltage of 10 millivolts or less in the electrode being tested. Since it is not uncommon for residual voltages in the order of 1 to 10 volts to be present in a large grounding system, accurate measurement or even detection of the test signal can be a major problem. This situation also exists when the measurements are made on transmission structures equipped with long counterpoises and/or overhead ground wires. Secondly, as the size of a grounding system increases, the zone of influence of the grounding system increases and it becomes more difficult to establish a return electrode that is completely isolated from the electrode being tested. For any facility other than remote substations or transmission line structures, the problem is usually compounded by connection of numerous external grounds such as water mains, gas pipes, communication circuits and low voltage distribution neutrals to the main ground. These connections can effectively extend the zone of influence of the tested ground considerably and in an unpredictable pattern. Another significant but less understood factor affecting fall-of-potential measurement is soil structure. When the soil structure is not uniform, the shape of the fall-of-potential curve can be noticeably affected, as can the correct potential probe location for determining the true resistance of the unknown electrode when the traverse follows the same direction as the remote current electrode. This soil structure dependence, originally described in [4], is shown in Figure 9.4. The curves of this figure give the required potential probe location for various two-layer soil structures. The effects of soil structure on fall-of-potential measurements are discussed extensively in [3,4]. Generally, the zone of influence around a ground electrode in two-layered soil which has a higher resistivity lower layer is greater than one in soil with a lower resistivity bottom layer. This effect can be best understood by considering the path of least resistance for current leaving the ground electrode. When the surface layer is more conductive than the bottom layer, ground currents will tend to flow out horizontally rather than penetrate the higher resistivity layer below. This effect can make accurate resistance measurements very difficult to obtain in areas where a ground electrode is buried in soil overlying bedrock or other high resistivity sublayers. When very low ground resistances are being measured, it is usually necessary to locate the remote current electrode at significantly greater distances from the electrode under test than with small, high resistance grounds. When the moveable potential probe is located along the same direction as the current probe, inductive coupling between the current and voltage leads may introduce a significant error in the measured apparent resistance of the electrode. Simply separating the voltage and current leads may not be sufficient since the mutual impedance is only affected to a small degree by separation distance. For example, increasing the separation between two parallel conductors from 1.5 to 75 m (5 to 250 feet) may only lower the mutual impedance between the two conductors by a factor of 0.5. The actual mutual impedance between the current and voltage leads depends on soil resistivity, frequency, and length of parallel exposure. For soil resistivities in the order of 100 ohm-meters and a 9-7 separation of 8 m (25 feet), the mutual impedance is in the order of 0.4 ohms per kilometer. Thus, a test current of 2 amperes could introduce a 0.08 volt error in a 100 m (330 feet) long potential lead. For an unknown resistance of 0.1 ohms, this would be 40% of the voltage being induced in the measured electrode. The effect of coupling between voltage and current leads is to slant the fall-of-potential curve upwards to the right, tending to obscure any levelling off which may occur as the potential probe leaves the influence of the measured electrode. Figure 9.5, derived from a computer study of an actual test illustrates the effect of potential traverse direction and mutual coupling between voltage and current leads on the fall-of-potential curve. Figure 9.4 Required Potential Probe Location in a Two-Layer Structure Figure 9.5 shows a potential profile (curve 1) taken orthogonally to the direction of the return current electrode. The resistance of the grounding system being tested has been determined analytically to be 0.42 ohms. This figure also shows the fall-of-potential curve taken in the same direction as the return current electrode, (curve 2) and the related curve modified to account for a mutual impedance of 0.4 ohms per kilometer between voltage and current leads. If a fall-of-potenti al curve is taken from a direction other than towards the return current electrode, a "good" curve will always be obtained. That is, the curve will always level off at some point. If, however, the return current electrode has not been located outside of the zone of influence of the measured electrode, the result may be significantly lower than the true resistance value. To ensure that this does not occur, the return current electrode must be located well beyond the distance of the "knee" in the measured fall-of-potential curve. One means of determining if the current electrode is sufficiently remote, is to relocate the current electrode approximately 10% beyond the original location and observe if the measured 9-8 potential changes. If the potential does not change, the return current electrode is adequately located. If an increase in potential is observed, the current electrode should be relocated even further away. There may be cases where it simply is not possible to obtain a return current electrode that is outside the zone of influence of a very extensive grounding system. In these cases, the use of computer techniques may be the only means of obtaining accurate results from the measurement data. The use of a grounding computer program to interpret fall-of-potential data is described in reference [3]. NEGLECTING EFFECTS OF COUPLING BETWEEN LEADS INCLUDING EFFECTS OF MUTUAL IMPEDANCE BETWEEN LEADS ----- © FALL OF POTENTIAL CURVES IN SAME DIRECTION AS C2 CURRENT ELECTRODE FALL OF POTENTIAL CURVE AT RIGHT ANGLES TO C2 CURRENT ELECTRODE DISTANCE ALONG PROFILE (meters) Figure 9.5 Fall of potential Profile Results of an Actual Test 9.2.6 OVERHEAD GROUND WIRES The presence of an overhead ground wire connected to a transmission line structure can considerably complicate transmission line structure ground resistance measurements. Using conventional instrumentation such as a Megger null-balance instrument, it is virtually impossible to determine the resistance of the local ground compared to the much lower resistance usually created by a large number of adjacent structures, interconnected by the overhead ground wire. The key to determining the resistance of a specific structure ground equipped with an overhead ground wire lies in the impedance versus frequency characteristics of the ground. A transmission line structure ground, by itself, will typically exhibit an almost constant impedance over a wide frequency range. Adjacent grounds, interconnected by the overhead ground wire, will exhibit an increasing impedance with frequency as the impedance of the overhead ground wire becomes significant with respect to the parallel resistance of the structure grounds. At high frequencies of say 20 kHz or above, the effects of adjacent structures should be almost completely removed from the impedance measurement. 9-9 At least one manufacturer has developed an instrument specifically designed to take advantage of this effect to measure the resistance of transmission line structure grounds [8]. Since this instrument uses a fixed frequency to carry out the impedance measurement, it is not possible to verify, by plotting an impedance versus frequency curve, that an asymptotic value has been reached. A voltmeter/ammeter test set up which employs a variable frequency power source along with a frequency selective voltmeter can be used to carry out a similar measurement. Although considerably less convenient than the single instrument referred to above, this method does permit a meaningful impedance versus frequency plot to be obtained. Since mutual impedance coupling effects will be very significant at high frequencies, great care must be taken when carrying out these measurements using either of the described methods. In general, both voltage and current leads should be kept as short as possible, allowing for the usual probe location considerations of the fall-of-potential method. In addition, the current and voltage leads should be oriented at 900 in order to allow the minimum amount of coupling. Tests carried out on a 138 kV transmission line equipped with an overhead ground wire using the variable frequency voltmeter/ammeter technique are described in Chapter 10. For these tests, the overhead ground wire was disconnected from several of the pole grounds in order to measure the structure ground alone as well as with the overhead ground wire connected. Similar tests on a 500 kV transmission line are also described in this same chapter. 9.3 GROUND RESISTANCE MEASUREMENT INSTRUMENTATION The instrumentation required to carry out accurate ground resistance measurements depends very significantly on the structure being measured, on soil conditions, and on ambient electrical noise levels at the measurement site. For small, high resistance ground electrodes such as ground rods, isolated transmission line structures or small substations, portable earth testers such as the Megger null-balance, or direct reading earth testers such as the Vibroground, are the most commonly used instruments [2]. As discussed in Section 9.2, the limitations of these instruments become apparent as the resistance being measured drops below 1 ohm and ambient electrical noise levels increase. Also, at low values of ground electrode impedance, the reactive component of a ground electrode impedance may be significant and a test instrument capable of measuring the reactive and resistive components may be desirable. 9.3.1 PORTABLE GROUND TESTERS There are several models of portable ground testers on the market. Most employ either a hand-cranked magneto or internal battery-powered inverter as a power source and some type of bridge circuit to simultaneously detect voltage and current and give a readout in ohms when the instrument is balanced. The most commonly used instruments of this type, the Megger hand-cranked and battery-powered units, give a 3 digit readout in ohms via the three decade dials used to null the instrument galvanometer. A range selector switch is used to select a full scale reading of 9.99, 99.9, 999, or 9990 ohms. The hand-cranked version of the Megger instrument, by generating a variable frequency ac signal, is somewhat less sensitive to stray ground currents than the battery-powered models. As with all instruments of these types, the resolution becomes very poor as the measured apparent impedance drops to 1 ohm and below. In spite of some manufacturer's claims to the contrary, measurements 9-10 below 1 ohm with most portable instruments should be considered as approximate at best. All instruments of this type can be used to carry out measurements by any one of the three methods described in Section 9.2. 9.3.2 VOLTMETER/AMMETER METHOD Measuring instruments employing a null-balance bridge have offered an advantage of drawing no current through the voltage sensing leads in the balanced condition. However, the availability of inexpensive, high impedance (greater than 1 megohm) direct reading digital voltmeters has made the voltmeter/ammeter technique very practical for ground resistance measurements. The main advantage this method offers over the use of portable ground testers is in the increased current that can be passed between the two current probes. The disadvantage is in the considerably increased cost, complexity and bulk of the measurement apparatus. A typical voltmeter/ammeter test set-up for measuring earth resistance is shown in Figure 9.6. The interface unit shown in Figure 9.6 contains a variable transformer as well as step-up and step-down transformers for accomodating a wide range of loop impedances. Such equipment is usually constructed by the user or purchased as a custom assembly to meet specific testing requirements. The primary disadvantage of this arrangement is the inability to distinguish between ambient power frequency signals and the test signal. The set-up is only suitable when the voltages induced in the earth by the test current significantly exceed the stray voltage levels. This may not always be possible, particularily with ground impedances in the range of 0.5 ohms or less. DIGITAL VOLTMETER AC POWER SOURCE. RETURN - CURRENT ELECTRODE POTENTIAL PROBE INTERFACE UNIT ADJUSTABLE AUTO TRANSFORMER ISOLATING TRANSFORMER Figure 9.6 AMMETER ELECTRODE TO BE MEASURED Voltmeter/Ammeter Test Set-Up A significant improvement in the usefulness of the circuit of Figure 9.6 can be obtained if the frequency of the power source used can be adjusted to 5 Hz or more away from the frequency of the ambient interference and a selective frequency voltmeter substituted for the wide bandwidth digital meter shown. A meter such as the Hewlett Packard 3581C is capable of resolving a 3 Hz bandwidth signal with excellent rejection of out-of-band signals. Thus, test signal voltages well below the level of ambient interference can readily be measured with an accuracy of typically 1%. A 60 Hz rejection filter can be used at the voltmeter input if required [7]. 9-11 A further refinement to the test configuration is introduced by the use of a variable frequency power oscillator in order to measure the ground impedance over a wide range of frequencies. This permits identification of the resistive and reactive components of ground impedance and can be of considerable value when measuring transmission line structure grounds which are connected to a continuous overhead ground wire. By increasing the frequency to a sufficiently high value, usually 20 kHz or higher, it may be possible to observe only the impedance of the ground to which the current lead is attached. Considerable care must be taken when taking high frequency measurements however, since inductive and capacitive coupling effects between current and voltage leads may be very significant. Figure 9.7 illustrates a complete test configuration for this type of measurement (see also Figures 10.4 and 10.5 of Chapter 10). ADJUSTABLE FREQUENCY POWER SOURCE frequemcv SELECTIVE VOLTMETER RETURN JCURRENT Electrode POTENTIAL PROBE INTERFACE UNIT ADJUSTABLE AUTOTRANSFORMER ISOLATING TRANSFORMER Figure 9.7 WIDE FREQUENCY BAND AMMETER ELECTRODE TO BE MEASURED Functional Schematic - High Current Frequency Selective Ground Resistance Test Set-Up 9.3.3 HIGH FREQUENCY PORTABLE INSTRUMENTS A portable instrument has been developed [8] specifically for measuring the ground resistance of transmission line structures equipped with overhead ground conductors but not equipped with continuous counterpoises. This instrument is fully electronic and generates a 25 kHz test current from an internal Ni-Cad battery. The resulting structure ground potential rise is measured by the instrument and an apparent resistance obtained which has eliminated the effect of the overhead ground wire. Since this method uses the fall-of-potential technique, the usual considerations of electrode spacing and lead coupling must be observed. 9.3.4 OTHER MEASUREMENT SYSTEMS In an attempt to simplify the determination of resistance and reactive components of low impedance grounds, a measurement technique has been developed [5] which utilizes a Fast Fourrier Transform digital signal analyser and a white noise power source to measure the impedance and phase angle over a wide frequency range. The technique is basically the same 9-12 as the voltmeter/ammeter technique described previously, except for the signal source and metering devices. Figure 9.8 shows the general configuration of the instrumentation used in the measurement procedure. The digital signal analyser is a Hewlett Packard Model 5420. The primary advantage of this technique over a conventional voltmeter/ammeter technique is in the ease with which a wide frequency range plot of impedances can be obtained, eliminating the need to extrapolate measurements taken at a limited number of frequencies to obtain a 60 Hz value. The main disadvantage is in the relatively high cost and complexity of the measurement apparatus. REMOTE POTENTIAL PROBE DUAL CHANHEL FFT DIGITAL SIGNAL ANALYZER GROUNDING Figure 9.8 General Instrumentation Arrangement for the Measurement of Ground Impedances Using a FFT Digital Analyser A portable instrument has been developed [6,15] for measurement of the resistive and reactive components of ground impedances using a phase sensitive detector. The unit is described in [6] and offers excellent immunity to high levels of ambient power frequency interference, even when measuring ground resistances in the order of 0.1 to 0.2 ohms. Two versions of the instrument are being offered: The DET1, a unit similar to the one described in the reference paper and the DET2, a less sophisticated instrument, capable only of resistance measurements. For transmission line structure measurement, the DET2 instrument is more than adequate. In some recent tests [16] carried out on 300 kV transmission line structures the DET2 instrument performed very well, measuring resistances as low as 1.5 ohms and as high as 200 ohms with excellent stability and resolution. A conventional Megger null balance instrument was used to carry out duplicate measurement in several cases and yielded nearly identical results for structure resistances above 20 ohms. For low structure resistances, in particular measurements on structures to which an overhead ground wire was attached, it was not possible to obtain a satisfactory reading with the null balance instrument due to high levels of ambient interference. An additional advantage of the digital instrument was in the reduced measurement time, being about 10% of the average time required to obtain a reading on the null balance instrument. 9.4 MEASUREMENT TECHNIQUES As with most field measurement requirements, ground resistance testing is subject to numerous possibilities for the introduction of errors, detected or undetected, in the measurement process. The development of a sufficient understanding of the measurement process, combined with a carefully developed measurement technique is essential if reliable results are to be obtained. Some of the most common sources of errors in fall-of-potential measurements are: 9-13 • Inadequate separation of the unknown and auxiliary current electrodes. m Operation of instrumentation at inadequate sensitivity levels. • Location of auxiliary current electrodes or potential probes in the vicinity of buried metallic structures. • Inductive coupling between voltage and current leads. • Excessively high current probe resistance which can lead to parasitic capacitance and resistance errors [6]. • Incorrect interpretation of fall-of-potential data. • Uncalibrated instrumentation. As has been stressed previously, the susceptibility of a resistance measurement to errors generally increases as the measured resistance decreases, particularly below one ohm. It is recommended, however, that similar procedures be followed for all measurements and, as experience is gained, short cuts or simplifications can be made where it is known that the accuracy of the results will not be significantly affected. Good preparation is essential to the accomplishment of a successful field measurement program. A standard check list of items required is very helpful to ensure that no unpleasant surprises occur at the test site. A typical list for a field measurement using the voltmeter/ammeter technique is shown in Table 9.1. Prior to the commencement of measurements, all available "as built" drawings relating to underground services should be obtained in order to check for the location of buried pipes, counterpoise, cables, ducts or other structures which could affect the resistance measurements. Proposed remote current probe and potential traverse locations should be marked on the site plans to assure their suitability. ITEM 1 2 3 4 5 6 7 8 9 10 1 12 13 14 15 16 17 18 19 20 Note “ Table 9.1 DESCRIPTION QUANTITY PORTABLE TESTER A.C. GENERATOR TEST LEAD SPOOLS THREE FOOT GROUND RODS FOUR POUND HAMMERS VICE GRIPS TWO-WAY RADIOS '-■HIGH CURRENT INTERFACE UNIT *BATTERY POWERED FREQUENCY SELECTIVE VOLTMETER *BATTERY POWERED DIGITAL VOLTMETERS TEST LEADS FOR METERS DATA SHEETS FIBERGLASS SURVEY TAPES BEARING COMPASS SMALL TOOL KIT RULER AND DIVIDERS SITE PLANS PROGRAMMABLE CALCULATOR GRAPH PAD CAMERA AMD FILMS 1 1 4 5 2 2 1 1 1 2 As required As required 2 1 As required As required As required t 1 As required In the absence of high ambient noise levels, these units may be replaced by portable ground testers such as Meggers or the like. Resistance Measurement Equipment Check List 9-14 Once a remote current probe location has been established, the resistance of the current loop should be checked by a simple two terminal resistance measurement. There is no unique loop resistance value above which the test configuration should be rejected, although loop resistances in excess of 1,000 ohms are generaly quite difficult to work with. When the voltmeter/ammeter method is being used with test currents in excess of 1 ampere, safety and power source considerations will dictate that loop resistances not exceed one or two hundred ohms. Values below 100 ohms would be preferable. When a suitably low auxiliary electrode resistance is not attainable with a single ground rod, an array of ground rods can be used. To be effective, the rods should be located at least 4.5 to 6 m (15 to 20 feet) apart and driven as deeply as possible. Obviously, low lying areas with noticeable ground water are ideal for locating a remote electrode and should be used where possible. As mentioned in Section 9.2.4, mutual coupling between voltage and current leads may introduce significant errors into measurements of low impedance ground electrodes. Table 9.2 shows the approximate amount of mutual impedance between two parallel conductors laid on top of the soil for various soil resistivities, spacings and lengths of parallel exposure. This table does not account for possible coupling between additional lengths of wire on reels, a factor which can be very significant if adequate care is not exercised in the placement of the current and voltage spools. If the expected value of resistance to be measured is in the order of the mutual impedance given in Table 9.2, the potential measuring apparatus should be located completely separate from the current source and, if possible, the potential profile should be taken at right angles to the direction of the auxiliary current electrode. It is worthwhile to note that measurements using a reversing dc power source and a synchronous detector are not affected by mutual impedances between current and voltage leads. In any case, transmission line structure resistance measurements are unlikely to be significantly affected by mutual impedances between the test leads. Leakage resistance and stray capacitance between test leads and ground are other effects that can influence the results of low value ground impedance measurements [6]. These effects are unlikely to be of significance in most transmission line structure ground measurements although they could become significant in high frequency measurements of structure grounds connected to overhead ground wires. In order to minimize these effects, it is essential that test leads be kept dry, particularly if the leads contain any splices; current and voltage probe resistances be kept low and the leads be separated as much as possible. It is essential that the measured values of apparent resistance be plotted as a fall-of-potential traverse is being conducted. Only in this way can anomalies in the measurement be detected and accounted for, and an assesment made as to the adequacy of the remote current electrode location. It is usually useful to take readings at relatively large potential probe intervals initially and then to take intermediate readings where an apparent knee or "flat spot" in the curve has been observed. A primary requirement of the measurements is to determine whether or not the remote current electrode has been located outsi de the zone of influence of the electrode being tested. For small grounding systems, locating the return current electrode at a distance of at least 10 times the largest dimension of the grounding system will usually assure this. When this distance is unpractical, the remoteness of the current electrode must be determined by verifying that a potential traverse in the direction of the current electrode produces a distinct zone of nearly constant readings near the midpoint between the two electrodes. Alternatively, a traverse in a direction opposite or at right angles to the direction of the current electrode (C2) should produce constant apparent resistivity readings within the distance at which C2 has been located. As a further test, if C2 is relocated somewhat beyond the original location, the constant apparent resistivity values should be unchanged. 9-15 SOIL RESISTIVITY (OHM-METERS) CONDUCTOR SEPARATION (METERS) Sz| MUTUAL IMPEDANCE AT 60 Hz OHMS/KM 0HMS/M1LE i 1 i 100 500 2500 1 100 500 2500 10 10 10 0.311 0.1(0 0.46 100 500 2500 50 50 50 0.22 0.28 0.34 100 500 2500 250 250 250 0.11 0.16 0.22 1 1 0.51 0.57 0.63 “ ~~ Table 9.2 i i i , 1 1 i 1 l i i i i l i l i i 0.82 0.92 1.01 0.55 0.64 0.74 0.35 0.45 0.55 0.17 0.26 0.35 Mutual Impedance Between Two Parallel Conductors on the Surface of the Earth When the voltmeter/ammeter technique is used with test currents of a few hundred milliamperes or more, hazardous voltages may be produced at the auxiliary current electrode and in the soil immediately surrounding the current electrode. Consequently, considerable caution should be exercised when tests are carried out under these circumstances. Two-way radios should be used to maintain communication between the person working at the remote end of the traverse and the operator of the power source. A "dead-man" switch should always be used to energize the test apparatus so that there is no possibility of accidentally leaving the C2 electrode energized. Measuring tapes used in resistance testing should always be nonmetallic for two reasons. Firstly, the presence of a long metallic path along the line of the traverse could affect potential readings. Secondly, a metal tape could transfer hazardous potentials from the vicinity of the auxiliary current electrode to a person remote from the electrode. Finally, the issue of instrument calibration must always be addressed. The environmental extremes often encountered in field work may affect instrumentation accuracy. Direct sunlight on a hot summer day or temperatures below 0 °C can easily submit test instruments to temperatures outside their operational limits. It is a good practice to carry calibrated resistances of 1 and 10 ohms that can be inserted in series with the Cl lead on the electrode being measured. If the PI lead is placed above the shunt, the apparent resistance should increase by exactly the value of shunt, provided that contact resistances between the current lead and the ground electrode being tested are sufficiently low. This calibration procedure is illustrated in Figure 9.9. 9.5 EARTH RESISTIVITY TESTS The object of earth resistivity tests related to grounding system design is to assist in the determination of an appropriate soil model which can be used to predict the effect of the underlying soil characteristics on the performance of the grounding system during ground faults. 9-16 An accurate determination of earth resistivity characteristics from test results can be a complex task. Measurement results are often subject to different interpretation because of the characteristically complex variations in earth composition and because surface electrical test methods are remote from the structures influencing the test results. CALIBRATED SHUNT UNKNOWN RESISTANCE TO BE MEASURED = (R +R) NOTE Figure 9.9 RETURN CURRENT ELECTRODE This cheak does not identify measurement errors due to mutual coupling between voltage and current leads. Field Calibration Check of Ground Resistance Measuring Equipment Usually, the electrical characteristics of the earth are sufficiently uniform over horizontal distances to permit the consideration of the soil beneath typical industrial sites or transmission line structures to be uniform over horizontal dimensions. Similarly, vertical variations in resistivity can often be described by one, or more frequently, two distinct layers of earth. In the two layer case, the interface between the two layers of earth is often found within a few meters of the surface. Although a two layer earth structure model can be a major simplification of the real situation, it has been shown [11,12] to be an excellent representation of complex earth structures for electrical grounding calculations. In carrying out resistivity measurements for the design of transmission line structure grounds, it should be kept in mind that the dimensions of structure grounds are generally small and, as such, will not be influenced to a great degree by very deep soil resistivities. This should be considered when resistivity tests are planned and measurement results interpreted [11]. Resistivity determination methods and the theory behind the most commonly used methods are discussed in detail in Chapter 4. In this chapter only galvanic resistivity measurement methods such as the Wenner array method or driven ground rod method will be discussed. 9.5.1 THE FOUR POINT METHOD The most widely used measurement configuration is a four-probe method developed by Dr. F. Wenner of the U. S. Bureau of Standards [13]. As shown in Figure 9.10(a), four uniformly spaced earth probes are inserted into the earth surface in a straight line. The outer pair of probes are used as current input probes and the inner pair as potential references. Using the Wenner geometry, the apparent measured resistivity is: 4713 R P 1 + 2a ' (9-12) 2a ---- /a2 + 4b2 /4a2 + 4b2 9-17 where P = apparent soil resistivity in ohm-meters. R = ratio of measured voltage to test current in ohms, a = uniform probe spacing, in meters, b = uniform probe penetration depth, in meters. Considering probe penetration depth to be a constant, Equation 9-12 effectively describes the variation in measured resistivity as a funtion of probe separation, a. Physically, the greater the probe spacing, the greater the volume of earth encompassed by the test current in its traverse from Cl to C2 and hence, the greater depth of earth involved in the measurement. The task of accurately relating the apparent resistivity measured by this procedure and the true resistivity at specific depths is complex but to a first approximation, the apparent resistivity which is measured at probe spacing a may be considered indicative of the average resistivity to a depth a. For probe spacings where a»b, the above formula can be simplified by neglecting b. The relationship then becomes: p = 2TraR 3) For probe spacings of approximately 10 times the electrode depth, the error introduced by the simplified formula is less than 1%. For a probe spacing equal to the electrode depth, the apparent resistance value obtained from the 'simplified formula is 40% lower than the correct result. Thus, probe depth must be accounted for if the measurement is made for short spacings. If the measured values of apparent resistivity vary significantly with probe spacing, a nonuniform soil structure is indicated. The development of an appropriate two-layer soil model from measured resistivity data is discussed in detail in Chapter 4. The computer program RESIST has been developed for this purpose. An important variation of the four probe method is the unequally spaced or Schlumberger-Palmer arrangement shown in Figure 9.10(b). This method overcomes a shortcoming of the Wenner method often encountered at large probe spacings where the magnitude of the potential between the potential probes is too small to give reliable measurements. By moving the potential leads closer to the outer current electrodes, the potential value is increased and the sensitivity limitations encountered using the Wenner method may be overcome. For large probe spacings, the apparent resistivity according to the Schlumberger-Palmer method is given as: p = TTRc(c+d)/d (S-'lk) where c = Spacing between voltage and current probes, d = Spacing between inner voltage probes. R = Measured apparent resistance. 9.5.2 DRIVEN GROUND ROD METHOD The driven ground rod method [9,14] may, in some cases, be preferred to the Wenner method because of the type of ground electrode to be installed. If the electrode resistance is heavily dependent on the resistance of an array of ground rods, then this method may be considered to be most representative of the actual design situation. It has been shown, however, that the driven ground rod method produces results that are consistent with the Wenner method 9-18 only in reasonably uniform soils [14]. In nonuniform soils, the appropriate interpretation of the driven ground rod method results may become an exceedingly complex task. FOUR TERMINAL TEST SET O Q D.lL CURRENT PROBE POTENTIAL PROBES (a) Figure 9.10 CURRENT PROBE c, it—la s' X / CURRENT PROBE POTENTIAL PROBES (b) WENNER C2 CURRENT PROBE SCHLUMBERGER Four Point Method of Earth Resistivity Measurement In using the driven ground rod method, the fall-of-potential method is applied to measure the resistance of a single driven rod. The apparent resistivity is then determined from the relation: 2'irLR (9-15) P l1 n —j— 8L i1 d where = Apparent soil resistivity to depth L (ohm-meters). L = Length of driven rod in contact with earth (meters), d = Diameter of rod (meters). R = Measured value of resistance (ohms). p Because of the shallow penetration that can be achieved in practical situations, the very localized measurement area and the inaccuracies encountered in two layer soil conditions [14], the driven ground rod method is only normally suitable for use in limited circumstances such as small transmission line structure grounds. In areas of difficult terrain, the driven ground rod method may be the only practical means of determining soil resistivity. 9.5.3 RESISTIVITY MEASUREMENT INSTRUMENTATION In order to carry out resistivity measurements by the Wenner or similar methods, it is necessary to provide a source of test current and a means of measuring the voltage between the voltage sensing probes in the test electrode array, or a means of determining the ratio of this voltage to the test current. Numerous instrumentation packages have been developed to carry out either of the above determinations. These packages vary considerably in portability, sensitivity, operating convenience and cost. According to a recent survey [2] of electrial utilities throughout the 9-19 world, the Megger null-balance and direct-ohm reading type instruments are by far the most widely used instruments for measuring earth resistivity. These instruments are both very well suited for field work, being small in size and weight and relatively easy to operate. The test current supplied by these and most portable instruments is relatively small however, typically in the order of 40 milliamperes. In areas of high ambient noise, very high or very low resistivity, and particularly at large probe spacings, sensitivity can be a problem and these instruments may not be suitable if accurate measurements are required. In order to overcome the sensitivity limitations of portable instrumentation, a direct measurement technique using a voltmeter, ammeter and an ac current source, rated at several hundred watts or more can be used. In order to avoid interference problems with ambient 60 Hz fields or ground currents, it is usually necessary to use a power source which operates 5 Hz or more away from 60 Hz, and to use a frequency selective voltmeter to measure the voltage between the voltage sensing points. The power source for such measurements may be a portable ac generator which has been adjusted to a suitable frequency and transformed to an appropriate source voltage. It is somewhat more convenient to use a variable frequency power oscillator, although this adds considerably to the cost and source power requirements of the system because of the low efficiency of the power amplifier. As well, there is normally no interest in carrying out resistivity measurements over a range of frequencies. Thus, the ease of frequency adjustment is of less value than in ground electrode impedance measurements where the full capabilities of the power oscillator can be utilized. The DET2 digital earth tester described in Section 9.3.4 is very convenient in carrying out resistivity measurements because of the speed with which readings can be obtained and the excellent resolution of the instrument at low apparent resistance readings. The primary limitation of the unit is in the maximum current probe resistance which cannot exceed 7300 ohms for both electrodes in series. This can present considerable difficulties if the surface resistivity is greater than 2000 to 2500 ohm-meters. 9.5.4 MEASUREMENT TECHNIQUES The measurement of earth resistivity using any of the instrumentation systems described previously is a relatively straightforward procedure. Obvious examples of good measurement practice such as ensuring good probe to soil contact, avoiding traverses adjacent to or across large buried metallic structures and maximum separation of current and voltage leads should always be observed. As probe spacing increases, the susceptibility of the measurements to errors increases. Portable instruments of the Megger category with range scales should always be operated at the maximum available sensitivity setting, and should not be relied on when the most significant digit is down to zero, ie. with two digit accuracy. When using Megger type instruments, it can be helpful to insert a milliammeter in series with the current leads, to monitor the applied test current. A drop in test current to 10% or less with the hand-cranked type of instrument can often indicate a poor current probe contact and a possibly suspect reading. If a low test current is observed, the probes should be checked and driven deeper or moistened in order to improve the probe contact resistance. At very close probe spacings however, the probe should be kept as short as possible and the soil left undisturbed if representative readings are to be obtained. Instrument sensitivity is less of a problem at close probe spacings however, so that low test currents may produce satisfactory results. 9-20 One of the most common faults in resistivity measurements is a failure to take measurements out to sufficiently large probe spacings. As a minimum, the total traverse should cover the dimensions of the grounding electrode under consideration. Ideally, the maximum probe spacing should exceed the dimensions of the electrode under consideration by extending the traverse to three times the electrode dimensions. At very large probe spacings, coupling between voltage and current leads and interference in the voltage sensing leads can introduce measurement errors. Voltage and current leads should be separated as much as possible and probe resistances kept as low as possible. Mutual coupling between voltage and current leads will not normally be a problem if the apparent resistance is greater than 0.5 ohms. Below this, caution must be exercised. A source of coupling between voltage and current leads that is often overlooked is the close proximity of test lead spools. Significant coupling can result when two spools are located close to each other with their centers on a common axis. The spools should always be separated as much as possible and oriented with the spool axes at right angles. The reliability of field measurements can be improved considerably if the apparent resistivity for each probe spacing is calculated and plotted as the measurements are being taken. In this way, discontinuities in the measurements can be observed immediately and the source of the discontinuity identified. Buried pipes, tanks, recently excavated and filled areas, and rock outcroppings are all examples of sources of discontinuities in resistivity data. Such anomalies should be identified and logged and a determination made as to whether or not the disturbance is entirely localized and can be ignored or should be taken into account in the developement of an appropriate soil model. An orthogonal traverse should always be taken to check the results of any resistivity traverse. If the same traverse center is used the results of the two traverses should track fairly closely. Large differences, of say 30% or more, between apparent resistivity values at the same probe spacing could indicate a measurement anomaly. Plotting both traverses on the same graph is extremely helpful in locating these. A standard data sheet should be prepared and used for all resistivity measurement data recording. The sheet should contain spaces to record details of the test site, test equiment, and weather conditions as well as all of the measurement data. It is also very advisable to accurately locate the traverse location on a detailed site plan at the time the measurements are made. Photographs can also be very helpful. The objective of data recording should be to ensure that a third party not having participated in the measurements, can carry out a detailed and accurate interpretation of the measurement results. 9.6 SAFETY PRECAUTIONS It should be strongly impressed on all test personnel that a lethal potential can exist between a station or transmission structure ground if a power system fault involving the grounding system under test occurs while the ground tests are being made. Since one of the objectives of tests on a grounding system is to establish the location of remote ground for both current and potential electrodes, the leads to these electrodes must be treated as though a potential could exist between test leads and any point on a station ground grid or transmission line structure ground. It is also possible that long test leads run parallel to a transmission line could become energized by mutual impedance coupling between a faulted phase conductor and the test leads. The magnitude of potentials transferred by test leads during power system fault conditions can be considerable, particularly in the case of a faulted transmission line structure where ground potential rises in excess of 10 kV would not be uncommon. 9-21 Obviously, the preceding discussion points to the necessity of caution when handling the test leads, and under no circumstances should the two hands or other parts of the body be allowed to complete the circuit between points of possible high potential difference. It is true that the chances are extremely remote that a fault will occur while test leads are being handled, but this possibility should not be discounted. Therefore, the use of insulating shoes, gloves, blankets and other protection devices are advised whenever measurements are carried out at an energized power station or transmission line. In all cases, safety procedures and practices adopted by the particular organization involved should be followed. REFERENCES 1 - IEEE Recommended Guide for Measuring Ground Resistance and Potential Gradients in the Earth. IEEE Standard 81-1962. 2 - M. Bouchard, F. Dawalibi, D. Mukhedkar, "Survey on Ground Resistance and Earth Resistivity Measurements". IEEE transactions on PAS, Vol. PAS-96, July/August 1977, pp. 1076 (Paper A 77 204-1). 3 - F. Dawalibi, D. Mukhedkar, "Resistance Measurement of Large Grounding Systems". IEEE Transactions on PAS, Vol, PAS-98, November/December 1979, pp. 2348-2354. 4 - F. Dawalibi, D. Mukhedkar, "Ground Electrode Resistance Measurements Nonuniform Soils". IEEE Transactions on PAS, Vol. PAS-93, January 1974, pp. 109-115. in 5 - I. D. Lu, R. M. Shier, "Application of a Digital Signal Analyzer to the Measurement of Power System Ground Impedances". IEEE Transactions on PAS, Vol. PAS-100, April 1981, pp. 1918-1922. 6 - P. H. Reynolds, D. S. Ironside, A. H. Silcocks, J. B. Williams, "A New Instrument for Measuring Ground Impedances", IEEE PES, Paper A79 080-3, IEEE PES Winter Meeting, New York, 1979. 7 - F. P. Zupa, J. F. Laidig, "A Practical Ground Potential Rise Technique for Power Stations". IEEE Transactions on PAS, Vol. PAS-99, January/February 1980, pp. 207-216. 8 - J. Ufermann, K. Jahn, "High Frequency Earth Resistance Measuring Instrument". Brown and Boveri Journal No. 3/67, pp. 132-135. 9 - IEEE Standard No. 81, "Recommended Guide for Measuring Earth Resistivity Ground Impedance and Earth Surface Potentials of a Grounding System". New Edition, to be published in 1982. 9-22 10- E. B. Curdts, "Some of the Fundamental Aspects of Ground Resistance Measurements" AIEE Transactions on PAS, Vol. 77, part I, 1958, p. 760. 11- F. Dawalibi, D. Mukhedkar, "Influence of Ground Rods on Grounding Grids", IEEE Transactions on PAS, Vol. PAS-98, November/December 1979, pp. 2089-2098. 12- G. F. Tagg, "Earth Resistances". George Newnes Ltd., London, 1964. 13- F. Wenner, "A Method of Measuring Resistivity", National Bureau of Standards, Scientific paper 12, No. S-258, 1916, p. 499. 14- C. J. Blattner, "Study of Driven Ground Rods and Four Point Soil Resistivity Tests", IEEE Transactions, Vol. PAS-101, No. 8, August 1982, pp. 2837-2850. 15- D, S. Ironside, "Some Recent Developments in Portable Ground Resistance Test Instruments", paper presented at the High-Voltage Power System Grounding Workshop Sponsored by EPRI, May 12-14, 1982, Atlanta, Georgia. 16- I. Simpson, "Alcan Smelters and Chemicals Kitimat Works Grounding Study", David Bensted and Associates Ltd., Final report, June 1982. 9-23 COMPARISON BETWEEN MEASUREMENTS AND COMPUTATIONS 10.1 GENERAL Excellent agreement has been obtained between experimental results measured using power system grounding models and calculated results based on recently developed analytical equations. [1,2,3]. Good agreement has also been obtained between the computed surge response of transmission line towers and the measured values on experimental scale models of transmission line structures [4,5,6]. There is however, little published information giving comparisons between calculated and measured results of tests carried out on actual transmission line installations. One reason for this is the difficulty in obtaining accurate soil resistivity information at a particular site, since this often involves a determination of resistivity variations with depth as well as over lateral distances. In addition, tests involving high frequency or impulse currents are difficult to implement and require careful interpretation. Despite the above difficulties, successful measurement programs for both power and surge frequencies have been reported in the literature [7,8,9,10]. The key to successful completion of such tests is in the careful planning of the test program and in the selection of the most suitable test equipment. In this chapter, two of these test programs will be described in detail. These test programs are: ® The Ontario-Hydro/American Electric Power (OH-AEP) 500 kV 765 kV transmission line tests [7]. • The Electricit§ de France (EDF) rocket-triggered lightning test program at St. Privat d'Allier [9,10]. The main subject of this chapter however, consists of two recently completed measurement programs conducted within the scope of this EPRI project. These tests will be described in detail and the measurements will be compared with the analytical results obtained from the computer programs developed under this EPRI project. These measurement programs are: • The Rochester Gas and Electric Corporation (RGE) 115 kV H-frame wood pole transmission line tests.• • The Tennessee Valley Authority (TVA) 500 kV transmission line staged-fault tests. 10-1 There are several objectives in describing these tests. One is a validation of the analytical derivations (and therefore, the corresponding computer programs) presented in this report. It is believed however, that this goal has been reached in the excellent agreement obtained between scale models and analytical models [2,3]. The next objective is to show that a "practical" agreement can also be obtained when the predictions are compared with actual installation results where the variables are not all rigidly controllable and/or arbitrarily selectable. Included in this chapter is a detailed description of the test activities, test equipment, measurement problems, and solutions. The preparation of the test data for the various computer programs and the interpretation of the computed results is also described. Hence, when used in conjunction with the previous Chapter 9 on field measurements, this chapter should serve as a guide to the successful implementation of transmission line grounding measurements and subsequent test data interpretation. 10.2 THE RGE 115 kV TRANSMISSION LINE MEASUREMENTS A section of a Rochester Gas and Electric 115 kV transmission line near Fillmore, New York, on Route 19A was selected as a site to carry out a series of measurements on H-frame wood poles which are representative of a large portion of the 115-230 kV transmission lines in existence in North America. One objective of the measurements was to observe and describe the conditions which are likely to exist during routine field tests on energized transmission line poles and further, to interpret and discuss the field results with respect to the calculations carried out using the computer programs RESIST and GTOWER (see Appendices B and C). The measurements involved earth resistivity tests and fall-of-potential impedance/resistance measurements on the pole grounding electrodes. The ground impedance measurements were carried out with and without the double overhead ground wires connected to the pole ground being measured. 10.2.1 TEST SITE The site of the measurements, near Fillmore New York, was an area of relatively flat terrain. The transmission line was located on an old railway right of way. The elevated road bed was readily identifiable, although it was heavily overgrown with blackberry bushes. A footpath down the center of the right of way provided the only access to some of the transmission line structures. The surface soil layer of the elevated roadbed consisted of a cinder or coke material. This was the only place in the area where this material was found and was assumed to have been used as fill for the railroad bed. The area on each side of the right of way was farmland. Route 19A ran parallel with the transmission line, approximately 35 m (115 ft) west. The site of the remote current electrode, 300 m (984 ft) north of structure 161, was at the side of a small pond (see Figure 10.1). The soil at that location was a very heavy wet clay. The area between the transmission line right of way and Route 19A where three of the resistivity traverses were taken, is a RGE pole yard. It was apparent that some minor excavation had taken place at some time which involved filling in the ditch alongside the former railway line and cutting into the elevated roadbed. Poles A and B of transmission line structure 161 were located on the edge of the pole yard where the roadbed appeared to have been cut away. 10-2 Route 19 A Pole Yard 115 kV H-Frame Wood Route 19 A RaiIway Structure No.------------ Figure 10.1 (a) Rochester Gas aid Electric Test Site 3-POLE STRUCTURE (str.163) Figure 10.2 115 (b) 2-POLE STRUCTURE kV Wood Structures 10-3 (str. I6l) Two types of structures were investigated. Structure 163 was a 3-pole structure with a steel guy wire and anchor connected to the overhead ground wire (Figure 10.2(a)). Structures 159, 160, 161 and 162 were H-frame (2 pole) structures as shown in Figure 10.2(b). In all cases except structure 163, the test lead was connected to a bared portion of the copper wire which ran up the pole from the ground rod buried adjacent to the pole, to the overhead ground wire (see Figure 10.3(a)). In the case of structure 163, this connection was made to the guy wire (Figure 10.3(b)) which was connected to the overhead ground wires via a cable, presumably under considerable tension, across the top of the two outer poles (see Figure 10.2(a)). (a) TEST LEADS CONNECTED TO VERTICAL GROUND VI RE Figure 10.3 (b) TEST LEADS CONNECTED TO GUY WIRE Current (Cl) and Potential (PI) Test Lead Connections At the time the measurements were taken, the soil surface was very wet from heavy rains of the previous day. 10-4 10.2.2 TEST EQUIPMENT Two types of instrumentation were used in the measurements carried out at the RGE test site: • A hand-cranked Megger (null balance instrument model ET3) was used to carry out the earth resistivity measurements and resistance measurements on one of the structures (structure 161) with the connection to the overhead ground wires open-circuited. ® A voltmeter/ammeter test system consisting of a California Instruments model 501T power source, a Hewlett-Packard 3581C frequency-selective voltmeter and a Fluke model 8050A digital multimeter, was used to measure the ground impedances with the overhead ground wires attached and the ground resistances of several transmission line poles when the pole grounds were isolated from the overhead ground wires. These instruments, shown in Figure 10.4, were also used to measure the ground impedances for various frequencies up to a value of 5000 Hz. The test power source was a porTable 2.5 kW generator which powered the 501T variable frequency power unit. The 501T provided the test current via an interface unit designed to provide test voltage control, circuit isolation and a convenient interface with the voltage and current measuring apparatus. A functional schematic of the test setup is shown in Figure 10.5. Figure 10.4 Test Equipment The wire used for the current and voltage test leads was No. 16 AWG type TEW. Four 1000 foot spools of wire were used in these tests. The portable wire reels used were specially designed to ease the frequent winding and unwinding operations necessary during the measurements. 10-5 Portab le Generator Selective Vo 1tmete r Return Current E ?ect rode Potential Probe Variable Frequency Generator Interface Unit Figure 10.5 Elect rode to be 'Measured Functional Schematic of the Test Set Up 10.2.3 EARTH RESISTIVITY TESTS A total of four resistivity traverses were carried out at the location shown in Figure 10.6. The measurement results are given in Appendix F, Tables F-l to F-4, and are also plotted in Figures 10.7 to 10.11. The selection of a good location to carry out a traverse was restricted considerably by the heavy undergrowth beneath the transmission line. One traverse (RGE 04) was carried out directly underneath the line between structures 160 and 161 where a footpath existed. It was not possible to carry out an orthogonal traverse at this location. Three other traverses were carried out in the pole yard. Two of these traverses (RGE 01-1 and RGE 03) gave inconsistent and erratic readings as probe spacings increased beyond 3 meters. The apparent resistivity values obtained in the two questionable traverses dropped very rapidly beyond the 3 m probe spacing. It was suggested by the RGE personnel present at the test site that the fill used when the pole yard was built could have contained discarded scrap metal. This would explain the results obtained. Traverses RGE 01-1 and RGE 01-2 were centered approximately 70 m (230 ft) north and 20 m (67 ft) west of structure 161. The first traverse, taken parallel to the transmission line, gave reasonable results for probe spacings of 0.5, 1 and 2 m but showed a sharp drop in apparent resistivity at spacings of 3 and 5 m. The results of the orthogonal traverse (RGE 01-2) were very similar to those of the first traverse at spacings of 2 m or less but, showed slightly increasing resistivity values at the probe spacings up to 10 m, rather than the sharp drop in apparent resistivity observed in the first traverse. It was concluded that buried metal structures extending along the orthogonal traverse line were responsible for the sharp decrease in the apparent resistivity of traverse RGE 01-1. Other effects, such as interference from the transmission line, could not possibly account for such large measurement differences. Traverse RGE 04-4 was located immediately underneath the transmission line and was centered between structures 160 and 161. A footpath along the right of way facilitated the measurements in this location. The results of this traverse differed from those of the previous pole yard traverse by a factor of approximately 1 to 2 for short spacings (except for a spacing of 0.5 m). This difference is attributed to the different surface covering described earlier and to the fact that the entire traverse was along an elevated path approximately 1.5 m above the level of the surrounding terrain. The values measured at close probe spacings indicate that the fill material constituting the elevated path has a low resistivity value. The apparent resistivity values approached the results of traverse RGE 02-1 as the probe spacing increased. 10-6 Return Current Electrode Fa t1-of-Poten tia Profiles T rave rse Traverse 03 01 ,2 I ravers' Traverse W? re Ok Route 19 A Route 19 A Pole A Structure Figure 10.6 St ructure St ructure St ructure St ructu re" Rochester Gas & Electric Measurement Site 115 kV Transmission Line 10.2.4 ANALYSIS OF THE RESISTIVITY TESTS The preceding qualitative analysis of the field results can not provide all the necessary information required to develop a suitable equivalent soil model for grounding studies. Computer program RESIST (Appendix B) was used to determine the equivalent soil models related to each resistivity traverse. Other methods described in Chapter 4 could be used also. In particular, the graphical chart provided in Volume 2 of this report will lead to similar results and is very convenient for a quick estimate or in the absence of computer facilities. RESIST determines a two-layer soil which is defined by three parameters; top-layer resistivity Pi, top layer thickness h and bottom layer resistivity P2. These are later used as input data to grounding computer programs such as GTOWER (Appendix C). When the measured apparent resistivity traverse exhibits a shape which indicates that the soil is too complex to be modelled by a two-layer structure, the resistivity traverse can be subdivided into appropriate segments which are then analyzed separately. Each traverse segment is then identified with a two-layer soil and the final equivalent two-layer soil model is constructed from engineering considerations, such as the extent and burial depth of the grounding system at the measurement site. This approach was used to analyze traverse RGE 04 which exhibits a three-layer earth structure with clearly different resistivity values for each layer. The analysis of traverses RGE 01-1, RGE 01-2 and RGE 03 using RESIST program was straightforward since the measured apparent resistivity curves in all three cases indicated a two-layer earth structure. The equivalent models derived by RESIST are given in Table 10.1. These models lead to an excellent curve match with the field results as shown in Figures 10.7, 10.8 and 10.9 respectively. 10-7 TRAVERSE EQUIVALENT SOIL MODEL h P2 Pi (m) (ft-m) (n-m) DESCRIPTION/COMMENTS RGE 01-1 Parallel to transmission line 207.6 14.4 1.41 0.75 RGE 01-2 Perpendicular to transmission line 209.9 110.3 RGE 01-3 Parallel to transmission line 165.1 38.3 2A RGE 04 Between structures 160 and 161 (0.5 m spacing ignored) 54.2 149.2 1.5 58.3 337.9 2.8 156.6 30.6 25.9 First section of curve (up to a RGE 04 (a) spacing of 16 m and 0.5 m spacing ignored). Second section of curve (with first RGE 04 (b) two-layers replaced by an equivalent sing1e-1ayer) Table 10.1 Equivalent Two-layer Soils APPARENT RESISTIVITY (ohm-meters) Traverse: rge 01-1 Computed Apparent Res Istivlty Measured Apparent Resistivity 160 - PROBE SPACING (meters) Figure 10.7 Earth Resistivity Traverse RGE 01-1 10-8 2k0 . Computed Apparent Resistivity .®~- Measured Apparent Resistivity Traverse:«« 01-2 PROBE SPACING (meters) Figure 10.8 Earth Resistivity Traverse RGE 01-2 Traverse: RGE 03 Computed Apparent Resistivity Measured Apparent Resistivity PROBE SPACING (meters) Figure 10.9 Earth Resistivity Traverse RGE 03 10-9 Traverse RGE 04 was first analyzed with all the measurement data except for the very short spacing value of 0.5 m. This spacing was discarded because it reflects the higher surface layer resistivity of the footpath and therefore, will not influence the resistance calculations of the pole ground rods. As can be seen from Figure 10.10, the equivalent two-layer soil determined by RESIST (see Table 10.1) does not match the measurement results well, except for the short probe-spacing zone of the curve. A second RESIST run was carried out using the data from traverse RGE 04 but without the 20 and 30 m probe spacing values. The equivalent soil model computed by RESIST is given in Table 10.1. This soil model leads to excellent agreement between the measured and calculated values for all probe spacings below 20 m (see Figure 10.10). However, it also suggests a 340 Q -m infinitely deep bottom layer. This of course does not agree with the data which show decreasing soil resistivity with depth. Hence, the 340 ft-m resistivity layer must have a finite thickness, and a third layer with a lower resistivity value is indicated. Traverse: rge _ _ _ _ _ _ RGE 0i( _____ RGE OMa) Computed Apparent Resistivity _ _ _ _ _ _ RGE 04(b) —©---Measured Apparent Resistivity 160 . PROBE SPACING (meters) Figure 10.10 Earth Resistivity Traverse RGE 04 10-10 04 In order to determine this third layer resistivity and the thickness of the 340 fi-m second layer, a third RESIST run was completed with the 20 and 30 m probe-spacing data and a soil resistivity value at 1 m spacing, of 150 £2-m. The 1 m, 150 £2-m values represent an equivalent uniform soil which replaces the original 58.3 fl-m / 337.9 fi-m / 2.8 m two-layer model determined in the second RESIST run. This equivalent single - layer can be determined empirically using the partial curve matching technique referred to in Chapter 4. However, since our final goal is to derive an equivalent two-layer soil from the more exact three-layer model, engineering judgment and visual interpolation was used to determine the equivalent single-layer resistivity. One of these considerations was that all previous traverses had an average apparent resistivity value in the order of 100 to 200 fi-m. The results of this third RESIST run are summarized in Table 10.1. Thus, the soil structure at this location can be described by the following parameters: Top Layer Resistivity Px Thickness hj : ft-m 2.8 m 53 ; Mid-layer Resistivity P2 Thickness h£ : 338 Q-m : 23 m Bottom Layer Resistivity P3 : 30 Q-m Since the pole ground rods are less than 2.5 m long (8 foot ground rods), the bottom layer resistivity will have little effect on the resistance of the rods. Therefore, it is concluded that for the purpose of ground rod resistance calculations, the equivalent two-layer soil to be selected for traverse RGE 04 is the one obtained when the measurements at 20 and 30 m are ignored (RGE 04(a)). 10.2.5 GROUND IMPEDANCE MEASUREMENTS Impedance measurements were carried out on the grounds of five consecutive transmission line structures of the Rochester Gas and Electric 115 kV line (structures 159 to 163). Measurements were taken at all 5 structures with the overhead ground wires solidly connected to all pole grounds. The vertical wire connecting the local pole ground to the overhead ground wires was cut at each pole of structure 160 and 161 (see Figure 10.3(a)) and the pole ground resistances were measured individually. The return current electrode C2 was located 300 m (984 ft) north of structure 161 (Figure 10.6). Three one meter ground rods were driven into very wet soil at the side of a small pond and connected via short jumper wires. The resistance of the return electrode was subsequently estimated from the measurement results to be about 21 ohms (see Appendix F, Table F-5). Initially, fall-of-potential measurements were attempted using the Megger instrument. It was found however that a satisfactory null could not be obtained. The instrument galvanometer swung randomly from extreme left to right as the instrument was cranked and the selector dials adjusted to obtain a null. This type of instrument behaviour is normally exhibited when the level of stray voltage appearing across the voltage terminals of the instrument exceeds the test signal by a large factor of approximately 25 to 1. Thus for an apparent impedance of 2 ohms, a stray voltage of 2 volts or more is likely to cause measurement difficulties. 10- 11 Actual experience could vary considerably according to the resistance of the return electrode, interfering frequency and type of null-balance instrument used. The above difficulty is quite common when the transmission line is energized and low impedances are being measured. There are techniques which can be used to overcome this problem. The selective voltmeter/ammeter technique used at the site is one suitable technique which also permits frequency sweeps during the impedance measurements. The functional schematic of the test setup is shown in Figure 10.5. The results of the fall-of-potential profiles taken to determine the transmission line pole ground impedances are given in Appendix F (Tables F-5 to F-14). The first fall-of-potential profile was taken from structure 161 with the Cl and PI electrodes clamped to the ground conductor at pole A of structure 161. As can be seen from the results (Table F-5), the potential readings rose very quickly to a stable value, indicating a very small zone of influence around the local ground rod of pole A. Once the fall-of-potential profile has been established, the potential probe was relocated at the 240 m location where apparent impedance readings were taken over a range of frequencies from 50 Hz to 5000 Hz (Table F-7). The impedance increased noticeably with frequency, indicating the presence of a significant inductive component in the apparent ground impedance. A discrepancy in the value measured at 1000 Hz for current levels of 1.0 and 2.0 amperes was attributed to overloading and distortion of the power amplifier output at the higher current value. Without relocating either of the remote probes, C2 or P2, the Cl and PI electrodes were located on the ground at poles 160 B and 159 B. It was assumed at this point that nearly identical readings would have been obtained if the other pole ground (for a given structure) was used as the ground tie point. Similarly, it was anticipated that readings at successive structures would be very similar since the local ground would be expected to contribute little to the overall measured ground impedance, except at very high frequencies. The results were somewhat different than expected, however. The 50 Hz readings taken at the first three structures ranged from 1.68 ohms to 2.35 ohms. Structures 162 and 163 (measured later) gave values of 3.85 ohms and 1.15 ohms. At 1000 Hz, the impedance readings ranged from 5.5 ohms (163) to 7.15 ohms (162 B), a considerably narrower range (Tables F-5 to F-12). It was noted during some of the tests that the test current did not remain constant, even though the source voltage did not appear to be unstable. After discussions with RGE personnel concerning the anomalies in the measurement results and an examination of the inconsistencies, it was concluded that the connection between the overhead ground wires and the vertical wire to the local ground was the most likely source of the problem. A resistance reading, taken with a digital multimeter, between the two-ground wires on poles 161 A and 161 B measured 89 ohms. Since the two overhead ground wires were almost certainly tied together at adjacent structures by the horizontal steel cross member or cable (at guyed structures), the problem is most probably caused by a poor connection to the local ground, as originally suspected. The lowest values of impedance were obtained at structure 163, a three pole guyed tower. The ground connection in this instance was made to the guy wire since no local grounds, other than the guy wire anchor, were installed. Being under considerable tension and connected directly to both overhead ground wires, this ground was believed to be very well tied electrically to the overhead ground wires (Figure 10.3(a)). The readings taken at structure 163 were consistently the lowest of all readings at all frequencies. Thus, at all other structures, it was assumed that a contact resistance of from 0.5 to almost 3 ohms existed between the overhead ground wires and the pole ground wires at structures 159 through 162. An approximate equivalent circuit of the ground connections at the base of each 10-12 structure is shown in Figure 10.11. This contact resistance, which may lead to significant error in ground impedance measurements, is typical when low voltages are used during the measurements [11]. Connections to Adjacent Towers Overhead Ground Wire to Vertical Wire Contact Resistance local / Ground Resistance Pole A Figure 10.11 Pole B Equivalent Circuit of Grounding Connections at Double-Pole Structures The connection from the overhead ground wire to the local ground was cut at poles 161 A, 161 B, 160 A and 160 B in order to measure the local ground resistance at each pole. Pole 161 A and 161 B grounds were measured with the Megger instrument which was not affected by stray fields while measuring the relatively high resistance local grounds at poles 161 A and 161 B. The grounds at poles 160 A and 160 B were measured using the voltmeter/ammeter test system. Measurements were taken at frequencies ranging from 50 Hz to 2000 Hz with a test current of 1.0 ampere. A slight drop in apparent impedance was observed. Mutual impedance effects between test leads accounts for this apparent decrease in electrode resistance. The test results are shown in Tables F-13 and F-14 of Appendix F. 10.2.6 ANALYSIS OF THE GROUND IMPEDANCE MEASUREMENTS Since no resistivity measurements could be conducted in the immediate vicinity of the transmission line structures and because of the large variations in the resistivity values observed from one site to the other, no attempt was made to assign to each structure any specific two-layer soil from the models derived in Section 10.2.4 and summarized in Table 10.1. Instead, the resistance of a typical pole ground rod buried in the soils described in Table 10.1, was calculated using the computer program GTOWER. The results obtained from GTOWER are summarized in Table 10.2. RGE practice is to equip each pole of the two-pole structure with a 8 foot, 5/8" diameter ground rod buried at an average depth of 1 foot. The ground rods at each pole of the structure are connected to the overhead ground wires via a vertical wire stapled to the pole. The one rod and two rod results shown in Table 10.2 simulate the measurements made with the vertical wire cut at each pole of the structure and with the two structure pole grounds connected via a jumper wire. 10-13 No RESISTIVITY TRAVERSE ID SOIL STRUCTURE Pi (ft-m) GROUND RESISTANCE (ohms) P2 ONE-ROD Q-m) TWO-RODS Un i form 100.0 - 2 RGE 01-1 207.6 14.4 1.41 8.9 4.8 3 RGE 01-2 200.9 110.3 0.75 45.9 25.3 h RGE 03 165.0 38.3 2.4 36.7 19-5 5 RGE 04 54.2 149.2 1.5 32.7 18.8 6 RGE 04 (a) 58.3 337.9 2.8 27.2 16.8 Table 10.2 - 38.3 21.3 Pole and Structure Ground Resistances When the computed ground resistances are compared with the measured values at structure 160 (33 and 40 £2 for pole resistances and 20 £2 for structure resistance; see Appendix F, Table F-13), reasonable agreement is obtained for both pole and structure ground resistances when the reference traverses used to perform the calculations are RGE 01-2, RGE 03 and RGE 04. There is poor agreement however between the pole ground resistances computed using the same resistivity models and the measured results for structure 161 given in Table F-14 of Appendix F (average measured ground rod resistance is 85 £2). Assuming the ground electrode was intact, this difference can only be explained by the presence of a relatively small volume of soil material (surrounding structure 161) with a resistivity value 2.5 to 3 times the average local soil resistivity. This illustrates well the effects of local soil resistivity anomalies on small grounding systems such as short ground rods. The apparent ground impedance of a structure, when equipped with a continuously grounded overhead ground wire, is frequency dependent and will vary from a low resistance value at very low frequencies, to a higher, predominantly resistive value at intermediate frequencies (25 to 50 kHz) as the resistance of the local ground become dominant it will exhibit a high reactive impedance at lightning or surge frequencies (200 to 1000 kHz). This behaviour is depicted in Figure 10.12. Zs = Rs+ja)X (a) Frequency = 0Hz Figure 10.12 (b) Frequency = 25 kHz (c) Frequency = 500kHz Frequency Dependence of Transmission Line Grounds 10-14 Figure 10.12 shows a ladder network where R represents the transmission line structure ground impedance (all assumed equal to an average value); Rs is the ground resistance of the structure at the point of observation; Z is the self impedance per span of the overhead ground wires and X is the inductance of the structure or vertical ground wires along the poles. At low and intermediate frequencies, the structure inductance is negligible and can be ignored. At 0 Hz, the per span impedance Z reduces to the per span dc resistance of the ground wires (r0). The impedance of the ladder network, as seen from the fault or test location (without the resistance of the local structure) is [11,12]: Z = (Z + /z2 + kRZ)/b (10-1) e Therefore the total impedance measured at the structure, including the structure ground resistance, is: Z Z (Z + /z2 + 4RZ) = —----------------------------t 4Z + Z + /z2 + 4RZ s (10-2) Table 10.3 gives the calculated results from Equation 10-2, based on the measured structures of the RGE 115 kV transmission line. The average structure ground resistance R is assumed to be approximately 30 ohms, while the tested structure ground resistance Rs is assumed to be 20, 30 or 50 ohms (according to the measurements R(160) = 20 Q and R(161) = 46 fi). The per span impedance Z of the two 7 No. 7 alumoweld ground wires were calculated using the computer program LIN PA described in Appendix A (span length = 125 m). Finally the downlead inductance X is estimated to be approximately 15 mH. Frequency (Hz) R (ohms) s (ohms) per Span Impedance (ohms) r jail 20 50 400 2000 30 30 30 (ohms) jt)Lt rt Measured Note 0.510 1.543 1.67 1 1.487 0.536 1.581 1.15 2 50 1.513 0.558 1.613 2.35 3 20 2.660 1.946 3-296 3.85 1 2.7H 2.128 3.447 3.12 2 50 2.745 2.289 3.574 4.25 3 20 5.391 3.937 6.676 9-3 1 5.584 4.728 7.317 8.15 2 7.808 30 30 30 0.257 0.213 0.358 1.548 0.769 7.155 5.660 50 30 1 Zt 1 Equivalent impedance Calculated (ohms) (ohms) 1.456 20 1000000 1 Zt I Zt z R 30 19.66 73.6 2285. 50 (1) According to measurements, Rg - 20 Q, 5.494 5.673 9.6 3 21.91 - - 29.46 10.09 31.14 - - 49.01 11.42 50.32 - - corresponds to structure 160. Value measured from pole B. (2) Value measured at structure 163. (3) According to measurements, Rs = 50 ^corresponds to structure 161. Value measured from pole A. Table 10.3 Frequency Response of Transmission Line Structures The calculated and measured values of the apparent impedance Zj- are shown in Figure 10.13. Examination of this figure indicates that the agreement between measured and calculated values is reasonable, particularly when the effect of the contact resistance is considered. If this contact resistance (estimated at .5 to 2 ohms) is eliminated by shifting down all the measured values by, say 0.8 $1, the average measured values track the average computed curve very well. 10-15 0 Figure 10.13 Measured and Computed Apparent Impedance 10.3 THE TVA 500 kV TRANSMISSION LINE MEASUREMENTS The Tennessee Valley Authority conducted a series of staged fault tests on their 52 km (32.2 miles) long 500 kV transmission line between Johnsonville and Cumberland steam generating stations. The faults were all single phase-to-ground and were made at tower 130, a few meters away from the special fault structure, shown in Figure 10.14, which was used to initiate them. The fault structure, shown again in Figure 10.15 while an arc was developing, is located about 10 km (6.2 miles) from Cumberland. Although the tests were carried out by TVA to achieve several goals, one of the main objectives of the tests was to determine the distribution of the fault current between the overhead ground wires and the transmission line towers and the earth potential around the faulted tower so that comparisons could be made with values calculated using computer programs PATHS and GTOWER. A program of resistivity and ground resistance measurements was undertaken at the test site as part of the 500 kV staged fault tests, to accurately determine the parameters required for calculating the fault currents and earth potentials. Since the two overhead alumoweld ground wires of the 500 kV transmission line are normally insulated from the supporting towers to provide a circuit for carrier communication [13], the initial staged fault tests were made while the ground wire insulators at tower 130 and at 20 other adjacent towers (10 on each side of tower 130) were short-circuited to simulate a transmission line with ground wires solidly connected to the towers. The final test was conducted with the transmission line operating under normal conditions, i.e., with the ground wires insulated from the towers. The ground wire insulators provided a 2.2 cm (0.87") air gap. Under dry conditions, this air gap will flash over for an average voltage of 15 kV. 10-16 PHASE C Figure 10.14 Test Site; Tower 130 and Fault Structure Figure 10.15 Arc Initiated at Fault Structure 10-17 The complete measurements conducted by Tennessee Valley Authority (TVA) and Safe Engineering Services (SES) were as follows: Resistivity Measurements (October 1981) • 4 traverses around tower 130 (by TVA). • 3 traverses around tower 130 (by SES). Resistance Measurement • Tower 1 to tower • Tower 125 to tower 135 (carried out by TVA, October 1981). • Tower 130 (several measurements by SES, October 1981). 166 (carried out by TVA between May and July 1970). Impedance Measurements (SES, October 1981) • At tower 130 with ground wire insulators short-circuited at tower 130 and 20 other adjacent towers (10 on each side of tower 130). Earth Surface Potential Profiles (SES, October 1981) • Around tower 130 with ground wires insulated from tower. • Around tower 130 with ground wire insulators short-circuited attower 130 and 20 other adjacent towers. Staged Fault Tests (TVA, October 1981) The following measurements were made three times with the ground wires electrically connected to tower 130 and 20 other adjacent towers, and once with the ground wires insulated from the towers. • Total fault current at the fault location (Ij-). • Total fault current in ground wires at the fault location (Ig). • Current in ground wires arriving at towers 125 to 129 and 131 to • Current in ground wires leaving tower 125 and tower 135. • Current in ground wires arriving at Johnsonville and Cumberlandstations. • Current in neutral of transformers at Johnsonville and Cumberland stations. • 135. Total fault current contribution from Johnsonville and Cumberland stations (I; Ic). • Ground potential rise of Johnsonville and Cumberland Station grounding grids. • Tower 130 ground potential rise measured with respect to two different remote points 90° apart (with respect to transmission line right of way. 10- 18 The staged fault test measurements are illustrated in Figure 10.16. The results of the staged fault current measurements which were carried out between October 27 and October 30, 1981 are presented in Section 10.3.10 where they are compared with the calculated values. The results of the resistivity, ground resistance/impedance and potential profile tests carried out on October 27 and October 28, 1981 are given in detail in Appendix G. They are analysed and discussed in Sections 10.3.5 and 10.3.6. All the measurements taken by SES were made using the selective voltmeter and ammeter method described in Section 10.2.2. No portable power generator was required for these tests since 120 volts ac power was available at the site. Tower Ground Potential Rise -'etc --etc Phase A Remote Soi 1 Grid Grid T owe r Fault Structure Tower 129 Remote Soi 1 CUMBERLAND J0HNS0NV1 LIE Figure 10.16 ' Ground Tower Wire 131 Staged Fault Tests Measurements 10.3.1 TEST SITE The fault site, near 500 kV tower 130, is a wooded area of irregular terrain as illustrated by Figure 10.17. The topography is characterized by numerous small gulleys with a small stream or ground water at the bottom of each. The transmission line right of way was traversed by three such gulleys between towers 130 and 129. Although not evident at the test site, it was apparent at nearby locations that the area is underlain by old, stratified bedrock which has been thrust up into innumerable small ridges and domes. The soil overburden appears to vary in thickness from a few feet to a hundred feet or more. No exposed bedrock was visible in the vicinity of the test site. Thus, it was not possible to estimate the overburden height in that area other than by interpretation of resistivity measurement data. At the time of the measurements, the ground was very wet as a result of heavy rains in the preceding few days. Vehicle traffic on the right of way created a considerable disturbance to the terrain near the test site, as deep ruts in the ground were created by the spinning wheels of the numerous TVA and other vehicles at the site. The weather was sunny for most of the tests however, with an ambient temperature in the range of 10 to 15 °c. 10- 19 The areas on each side of the 500 kV right of way were heavily wooded. Vegetation on the right of way itself had recently been cut down to about 15 cm (61'). Portions of the right of way close to the fault location traversed open fields. Right-of-way 6.9 kV Tower (b) Gravel Road Top View of the Test Site Soi] Profile Center Profile 50 Ft Left ------ 50 Ft Right G rave 3 Road Second Gulley Fi rst Gu1 ley Profile View of the Test Site Figure 10.17 Topography ot the Test Site 10.3.2 500 kV STAGED FAULT TEST DESCRIPTION To create a fault on the 500 kV line, a drop wire from one phase of the 500 kV line was connected to a conductor strung between two wooden poles on opposite sides of the transmission line right of way. This cross conductor was placed approximately 15 meters (50 feet) above the ground as shown in Figure 10.14. A second conductor was strung approximately 5 meters (16.4 feet) below the 500 kV cross-conductor. This conductor was connected solidly to the metal structure of tower 130. 10-20 The fault was created by the vertical firing of an arrow, the end of which was attached to a long nylon wire terminated by a short length of a metallic wire which initiated the arc when passing by the cross-conductors. The arrow was shot from a specially modified crossbow which was mounted vertically in a wooden support stand (Figure 10.18). The crossbow was fired remotely via a solenoid-inverter arrangement which allowed the person activating the device to be standing well away from the faulted tower. TOWER 130 POTENTIAL TRANSFORMERS CROSSBOW Figure 10.18 Potential Transformers and Crossbow Near Tower 130 Once initiated, the arc developed into a large fireball between the parallel phase and ground cross-conductors and generated a loud thunder-like bang. Although very difficult to estimate, the duration of the fireball appeared to be approximately 1/4 to 1/2 second. The actual fault clearing time was apparently 3 to 4 cycles. The phase and ground wire currents were measured using suitably rated conventional current transformers. The secondary winding currents were locally coded, then transmitted to the measuring equipment trailer via fiber optic cables. There, the received signal was amplified, decoded and sent to the recording apparatus. This technique electrically isolated interference picked up from the energized transmission line. All the fiber optic cables and associated circuitry were tested and calibrated by TVA personnel a few days before the initiation of the first staged fault tests. Immediately after the first staged fault test on Tuesday October 27, it was apparent that something was wrong. Smoke was observed rising from the measuring equipment installed in a trailer at the test site. 10-21 Smoke was also observed at one of the two high-voltage potential transformers mounted on a small open trailer located immediately adjacent to the faulted tower (Figure 10.18). An inspection revealed that the faulted tower structure had been inadvertently tied to the potential transformer casings. This resulted in a flashover from the PT casings to the low voltage secondary windings which lead into the equipment trailer. Thus, the full potential rise of the tower (more than 10 kV) was transferred into the equipment trailer on low voltage leads. As a result, several circuit cards in the recording amplifiers were damaged. This test was partially successful in that some data were obtained prior to the equipment damage and measurements at the two stations at each end of the faulted line were successful. On the same day, a second test was carried out with the potential transformer ground lead problem rectified. Unfortunately, it was then learned that several of the optical fiber cables used to transmit the fault current data from adjacent towers had been deliberately cut by vandals. Out of a possible 14 overhead conductor current readings, only 6 were recorded. Specially trained personnel with the appropriate splicing kits were required to repair the damaged fiber optic cables. This resulted in a two day delay in completion of the tests. The final tests were carried out onFriday with no improvement over test No. 2 because, not only the repaired fiber optic cables did not perform correctly, but also the current readings from two additional towers (127 and 128) were lost, presumably because of new damaged sections of the corresponding fiber optic cables. Fortunately however, the most important readings related to transmission line grounding analysis were successfully recorded. These readings included the total fault current in the ground wires and the potential rise of the faulted tower 130. 10.3.3 LOW CURRENT TESTS Soil resistivity tower footing resistance and earth potential gradients were measured at the TVA staged fault test site to provide the best possible input data to the transmission line grounding programs used to analyze the faulted power system. All measurements carried out by SES used the voltmeter/ammeter method with test currents of up to 2 amperes. A 250 VA variable frequency power source (California Instruments model 251T) was used to supply the test current and a Hewlett Packard model 3581C selective frequency voltmeter was used to measure voltage. The test current was measured by a Fluke true rms digital voltmeter model 8010A. A schematic of the test instrumentation arrangement is shown in Figure 10.5. TVA personnel duplicated the resistivity and the ground resistance measurements (ground wires insulated) at tower 130 to compare measurements conducted using relatively sophisticated equipment against the more expedient type of measurement usually conducted by utilities using conventional equipment. In addition, TVA personnel repeated the ground resistance measurements at towers 125 to 135 (towers 1 to 168 were measured in 1970). These results are contained in Appendix G (Tables G-17 to G-19). The resistivity measurements were carried out along four orthogonal directions as shown in Figure 10.19(a). Four sets of test probes extending in these four directions from tower 130 were installed by TVA personnel prior to the SES tests. Each test probe consisted of a 1.2 m (4 foot) ground rod driven to a depth of 0.9 m (3 feet). These probes were used for resistivity and fall-of-potential measurements. Additional temporary probes were used to measure potential gradients and resistivity for close probe spacings underneath and immediately adjacent to tower 130. The resistivity measurements for close probe spacings were repeated using the short, temporary probes instead of the 0.9 m (3 foot) probes. Six days prior to the SES tests, TVA personnel had carried out resistivity measurements with a Biddle Megger 10-22 model 597 instrument using the probe locations identified on the site plan. The results of these measurements are contained in Appendix G (Tables G-l to G-7). All resistivity measurements were based on the Wenner array configuration shown in Figure 10.19(b). The potential profile measurements, which were also used to determine fall-of-potential ground impedance measurements, were labelled as indicated in Figure 10.25 and Figure 10.26. The measurements were made with the ground wires insulated and short-circuited as already explained. The potential results of the profile measurements are given in Appendix G (Tables G-8 to G-16). Traverse 03 and Traverse 2- Buried Leg Anchor Traverse 10 and Traverse f-2\ Traverse Traverse 04 Traverse 2-2 b a JOHNSONVILLE CUMBERLAND (a) Location of Trave Figure 10.19 (b) Wenner Method Resistivity Measurements 10.3.4 SES TEST PROCEDURE The same basic procedure was applied in all of the measurements carried out by SES. Once the four test leads (two current and two voltage leads) had been located, an ac voltage of up to 160 volts was applied between the current electrodes from the power oscillator. The multimeter was located in series with the low voltage side of the power oscillator output. In all tests, the output current was set to as high a value as could be obtained without exceeding the 250 VA power output or voltage limits on the power oscillator. Since the rating of the oscillator was limited at very low or very high frequencies, some low frequency measurements had to be carried out at reduced power levels. The primary factor determining output current, however, was the resistance of the current electrodes. For the resistance and potential profile tests, test currents in the order of 1.5 amperes were used. In order to achieve this current level, it was necessary to drive multiple ground rods interconnected via a horizontally buried bare copper wire. The test currents used for resistivity measurements were considerably lower because of the high resistance of a single rod and the 160 volt limit on the power source. Test currents for resistivity ranged from as low as 0.01 amperes to 0.20 amperes. 10-23 The nominal test frequency used in all tests was 70 Hz. When a test at a different frequency was required, the test probe locations were left fixed while test currents of different frequencies were applied and the appropriate voltage readings taken. Once a stable test current was obtained, the voltage between the two potential leads was measured using the selective voltmeter. This instrument was used at the 3 Hz bandwidth switch position to ensure rejection of all 60 Hz ambient noise and the AFC feature was used to ensure that the receiver remained tuned precisely on the power source frequency. As a calibration check, a Beckman 3020 digital multimeter was periodically connected in parallel with the selective voltmeter. Although the Beckman is a wideband meter, for many of the measurements the out of band signal voltages were a factor of 10 or more below the measured test voltage and thus had negligible influence on the measured wideband voltage. The selective voltmeter was also calibrated frequently using the instrument's internal calibration feature. 10.3.5 EARTH RESISTIVITY TESTS The resistivity traverses which were carried out by TVA and SES are illustrated in Figure 10.19(a). The measured results for these traverses are given in Appendix G and plotted in Figures 10.20 to 10.24. Two SES traverses, TVA 03 and TVA 10, utilized the same installed ground rods as used by TVA in tests carried out the week before. All resistivity measurements were based on the Wenner array configuration shown in Figure 10.19(b). Traverse: .... TVA 1- Computed by RESIST Measured by TVA (Megger) 240 _ This measurement point was discarded during the computations 160 _ Equivalent Soil 1.1m (3.6 feet) PROBE SPACING Figure 10.20 Resistivity Traverse TVA 1-1 10-24 (feet) 360 Traverse: tva 2-2 200- Computed by RESIST Measured by TVA (Meggi Equivalent Sol Pi = 31^ 97 h . 20 ■ i b0 ■ 1 60 ■ = 3.1?m (10.A feet r---------- *---------- 1-----------'---------- j-----------'---------- r 100 80 PROBE SPACING Figure 10.21 120 140 (feet) Resistivity Traverse TVA 2-2 Because of the major effects that probe depth can have on resistivity measurements with close probe spacings, the measurement data for 0.45 m (1.5 foot) probe spacing were analyzed using an accurate form of the apparent resistivity formula (see Appendices F or G). This formula gives the "average" surface layer resistivity to a depth approximately equal to the probe spacing. During the field resistivity tests the temporary probes were driven a few centimeters into the soil for 0.45 m (1.5 foot) probe spacing. Resistivities in the order of 600 to 1000 ohm-meters were recorded. This is about two to five times the values measured using the 0.9 meter (3 foot) preinstalled ground rods. The high surface soil resistivity is caused by the presence of small trees and the considerable quantity of other vegetation which had been recently cut a few centimeters from the ground (Figure 10.27). Because the surface soil resistivity has a negligible effect on the ground resistance and the surface potentials (at an average depth exceeding 8 centimeters), the calculations should be performed based on the "average" surface resistivity measured with the deeper preinstalled ground rods. 10-25 360 . Traverse:. TVA 280 . 1-2 200 . Computed by RESIST TVA 1-2 Measured by TVA (Megger) ------- Computed by RESIST (partial Curve Matching Technique) TVA 10 • Measured by SES (voltmeter/ammeter) ® Not included in computations (both sections of curve) E Not included in computations (first section of curve) EQUIVALENT SOIL (TVA 10) First section of curve EQUIVALENT SOIL Last section of curve Pi - 352 fi-m P2 - 162.5 ^“m hi = l.kl m (8.1') PROBE SPACING Figure 10.22 (TVA 1-2) Pi = 309 ft-m P2 = 209 ft-m h = 1.35 m (4. A (feet) Resistivity Traverses TVA 1-2 and TVA 10 The measurement results show noticeable differences between the various traverses. However, there are several similarities and a general trend of the apparent resistivity curves. All curves indicate a complex earth structure, which, considering the nature of the terrain, is to be expected. Examination of Figures 10.20 to 10.24 and the following analysis indicates that the soil structure around tower 130 should be treated with a three to four layer model, with a thin moderate resistivity top-layer (P} = 250 to 350 P-m, hi = 1.5 m) underlain by a relatively thick low resistivity layer (P2 = 70 to 90 P-m, h2 = 4 m). These two upper-layers rest on one or more soil layers with an average resistivity in the order of 400 P-m (a typical value for ancient layered-rock formations). Since it is not yet practical to analyze grounding problems in soils with more than two layers, it is necessary to find a suitable equivalent two-layer soil model for the real soil conditions observed at the test site. Unfortunately, the problem is further complicated because, even within a small radius from the tower, the earth structure appears to vary with the direction of the traverse being examined. Therefore, to arrive at a "practical" equivalent soil model which will represent the tower grounding system with reasonable accuracy, averaging techniques must be employed on all traverses as well as sound engineering judgment. 10-26 Fortunately, in most power grounding problems, significant variations in the soil characteristics can be tolerated at moderate or large distances from the ground electrode, without noticeable effects on its grounding performance. By moderate distances it is meant about two to three times the largest dimension of the grounding system. It is therefore permissible and quite accurate to replace the "distant" soil layers by an equivalent one. In many instances this equivalencing (or averaging) technique can even be performed at relatively short distances from the ground conductors without significantly changing the essential aspects of the grounding system response to fault currents. Such transformation techniques are often applied in geological prospecting and use the concept of transversal and longitudinal earth resistivity [14]. This concept however can not always be applied in power grounding because of the fundamental difference between a point source ground electrode which is in contact with one single soil layer at a time, and an extensive grounding system which may be in contact with more than one layer and hence, be poorly modelled if two of these layers are replaced with one single equivalent layer. 320 -| . TVA 203 Traverse: TVA --------- Computed by RESIST ....... Measured by TVA (Megger) _ ____ Computed by RESIST TVA 03 ...4.-.- Measured by SES (vol tmeter/ammeter) 200 . 160 - 120 _ EQUIVALENT SOIL TVA 03 Pi = 3^ h = 0.8Am (2.75 feet) h = 0.8m (2.6 feet) PROBE SPACING Figure 10.23 (feet) Resistivity Traverses TVA 2-1 and TVA 03 10-27 320 _ Traverse: tva oa __ Computed by RESIST (Partial Curve Matching Technique) Measured by SES (vol tmeter/ammeter) Equivalent EQUIVALENT SOIL FIRST SECTION OF CURVE LAST SECTION OF CURVE Pi = 280.5 P 3 = 321. • 93 rn (29.3 feet PROBE SPACING Figure 10.24 (feet) Resistivity Traverse TVA 04 Consequently, the following principle must always guide the power engineer during the development of an equivalent or average soil structure: • Always combine layers which are not in direct contact with major parts of a grounding system. Try to maintain the integrity of the real earth structure where soil is in contact with the ground conductors. This principle has been followed below. However, in other situations it may be impossible to apply. In such cases, several different two-layer soil models encompassing the real model should be constructed and analyzed as the limiting cases. 10.3.6 ANALYSIS OF THE RESISTIVITY TESTS All resistivity traverses were analyzed using computer program RESIST to derive an equivalent two-layer soil model from the apparent resistivity measurements, for use in the tower grounding analysis. The results of this analysis are plotted in Figures 10.20 to 10.24 and are summarized in Table 10.4. In four cases (TVA 1-1, 2-1, 03 and 2-2) a reasonable equivalent two-layer soil was obtained for probe spacing of 25 m (80 ft) and more. In three other cases however (TVA 1-2, 10 and 04) the presence of the deeper layers was noticeable at shorter probe spacings and therefore, a partial curve matching technique was necessary to distinguish the predominant soil layers. 10-28 The results of the analyses are displayed in Figures 10.22 and 10.24. The two-layer equivalent soil models which were derived in each case by keeping the first layer resistivity (pj) and height (hj) unchanged and replacing the mid and bottom layers with an equivalent one, are given in Table 10.4. This derivation, which is not unique, is largely empirical and based on experience and judgment. It will be shown later that other values (in the same order of magnitude) would lead to small but noticeable changes in the actual computed values. During a design stage, such parameter variations should be used to determine worst case situations. In a comparative study such as this, the uncertainties involved in the proper choice of the parameter value suggest that a perfect match with the measured data will be purely coincidental. The closeness of the agreement will therefore be limited by the computed difference between the two probable, but reasonable extreme values of the uncertain parameters. RESISTIVITY TRAVERSE EQUIVALENT SOIL MODEL DESCR1PT1 ON/COMMENTS Pi (a-m) h (m) P2 (Q-m) MEASURED BY TVA 1-1 Along transmission line (toward Cumberland) 250 130 1.1 TVA TVA 2-2 Perpendicular to transmission 1ine (south East) vk 197 3.2 TVA 309 203 1.35 TVA 352 162.5* 2.47 SES TVA 1-2 TVA 10 TVA 2-1 TVA 03 TVA Qi) Along transmission line (toward Johnsonvi1le) Perpendicular to transmission 1 i ne (north-wes t) Along transmission line (centered at tower 130) 310 118 0.8 TVA ikh 113 0.84 SES 280 120 2.56 Note * Equivalent single layer Table 10.4 Equivalent Two-layer Soils The SES resistivity curves (measured using the selective voltmeter/ammeter method) and the TVA resistivity curves (measured using a Megger) agree quite well although the SES measurements give consistently slightly higher values. Resistivity traverse 04 was centered underneath tower 130 and extended outwards uniformly in each direction underneath and parallel to the 500 kV line. This traverse utilized the installed 0.9 m (3 foot) ground rods on both sides of the tower. At close probe spacings, temporary short rods were driven at appropriate locations. This arrangement permitted a traverse with 46 m (150 foot) maximum probe spacing. Test currents of 0.025 amperes were used for close probe spacings while tests currents of up to 0.2 amperes were used at large probe spacings. Note that the use of very low test currents (such as those generated by a Megger) might not have been practical had the transmission line been in service. However this would have depended on the degree of line imbalance, test lead arrangement with respect to the line, soil resistivity, ground rod depth, and other factors. 10-29 The resistivity tests and subsequent interpretation analyses suggest several equivalent two-layer soil models characterized by the following ranges of values (see Table 10.4): P1 = 250 to 350 tt-m P 2 = 110 to 210 Q -m h = 1 to 3 m These variations are not due to measurement uncertainties, and represent the real situation at the tower site. Despite the variations, the various models remain equivalent. For example, when comparing two models, pj increases from 250 to 344 fi-m, P2 decreases from 130 to 113 SI -m and h drops from 1.1 to 0.84 m. These variations have opposite actions and tend to cancel so that the various models will produce comparable results. Therefore, the potential profile and ground resistance analysis of Section 10.3.6 could be based on all the equivalent soils derived from the traverses (Table 10.4). However, to simulate a normal test program and to establish a fair basis for the comparisons, only one of the duplicate measurements (made by SES and TVA) for the same traverse should be considered. It was therefore decided to analyze both measurements separately and to use the SES based computation results as the reference for comparison with the measured earth potential and ground resistance values, since these were obtained using the same test equipment. Finally the sensitivity of tower grounding performance to small variations in the soil characteristics was established by considering additional soil models, qualitatively (logically) and quantitatively compatible with the measurement results. A total of ten different soil models are examined in the next section. Table 10.5 describes these models and provides a brief explanation of the methodology used to construct them. CHARACTERISTICS OF MODEL h Model No. (fi-m) (^-m) i 250 130 1.1 3.6 2 31^ 197 3.2 10.5 3 3^4 113 0.84 h 352 162.5 5 280 6 (m) DESCRIPTION OF MODEL CONSTRUCTION METHODOLOGY REFERENCE (feet) Derived from RESIST computations (traverse TVA 1-1) TVA measured Derived from RESIST computations (traverse TVA 2-2) TVA measured 2.75 Derived from RESIST computations (traverse TVA 03) TVA measured 2.47 8.1 Derived from RESIST computations (traverse TVA 10) SES measured 120 2.56 8.4 Derived from RESIST computations (traverse TVA 04) SES measured 308 ISO 1.76 5.8 Averaqe of all values in table 10-4 TVA + SES averaged 7 296 163.5 1.61 5.3 Averaqe of TVA values in table 10-4 TVA averaged 8* 325 132 1.95 6.4 Averaqe of SES values in table 10-4 SES averaged 9 300 143 1.4 4.6 Compatible Model (could have been compatible selected by other investigators) 10 341 132 1.52 5.0 Compatible Model (could have been compatible selected by other investigators) Note * Reference Soil Model for Grounding Computations. Table 10.5 Origin of Soil Models Analysed 10-30 10.3.7 EARTH POTENTIAL AND GROUND IMPEDANCE TESTS The potential measurements at 500 kV tower 130 were taken in two stages. First, measurements were carried out with the overhead ground wires connected to the test tower 130. The C2 return current electrode consisted of three vertical rods interconnected by a length of bare copper wire which was buried 5 to 8 cm (2-3 in) below the mud at the bottom of a small gully, 95 m (315 feet) west of the test tower, along the transmission line right of way (Figure 10.25). This electrode was determined to have a resistance of 95.5 ohms. The first fall-of-potential profile TVA 01 was taken at right angles to the transmission line (north-west) and utilized the ground rods previously installed by TVA personnel. Potential readings were taken out to a distance of 65 m (215 feet). With the potential probe at the 65 m location, the tower current electrode Cl was located on all four corner legs of tower 130. There was no measurable difference in readings. The test frequency was then varied from 20 Hz to 20 kHz and the apparent tower impedance determined for each frequency. The second fall-of-potential profile, TVA 02, was in the same direction as the return current electrode and again used the previously installed ground rods as potential measurement probes. In this profile, as well as the previous one, the first measured point (0 m) was at the center of the tower. Thus, the 0 m reading is not zero volts because the tower is in contact with the earth at the corners only. Immediately following profile TVA 02, a detailed potential profile (TVA 11) was measured between the south and west tower legs. The results of all these potential profiles are given in Appendix G (Tables G-8 to G-ll). Measuring Equipment Profile 01 95 m (315 Potentia 1 Current Lead (Cl) Current Profile 02 Return Current E1ect rode (In first Gu1 ley) Potentia 1 Probe P Tower Leg JOHNSONVILLE Figure 10.25 Buried Leg Anchor Profi1e 11 CUMBERLAND Potential Profile Measurements (Ground Wires Connected to Tower 130 and 10 Adjacent Towers on Each Side of Tower 130) 10-31 For the next series of potential measurements taken the following day, the overhead ground wire was insulated from the tower. This was the normal operating condition for this line. The first traverse (TVA 05), shown in Figure 10.26, was in the opposite direction to the current electrode. The current electrode (C2b) was initially the same electrode as used for TVA 01 and TVA 02 but was then relocated 180 m (600 feet) further down the transmission line right of way. The potential reading at 96 m (315 feet) changed by 1.2% when the current electrode was relocated. The new C2b electrode was higher in resistance than the first, therefore a lower test current had to be used. Potential profile TVA 05 used the installed ground rods as potential probe locations. Potential profile TVA 06 used the same C2b electrode as TVA 05 and was taken towards this electrode. Thus, TVA 06 was a repeat of TVA 02 but with the overhead ground wires insulated from the tower. Similarly, TVA 07 was a repeat of TVA 01 with the overhead ground wires isolated. Return Current Electrode n Second Gulley • C2b Profile 07 Prof lie 09 In First Gulley Potential Probe / Profile 08 Leg to Leg) # C2a Profile 06 Profile 05 Buried Anchor Tower Leg 1 80 m CUMBERLAND JOHNSONVILLE Figure 10.26 Potential Profile Measurements (Ground Wires Disconnected From Tower 130) Profile TVA 08 was a very detailed potential profile between the north and west tower legs of tower 130 (see Figure 10.27). The final potential profile TVA 09, was taken in a south-east direction from tower 130, once again using the ground rods installed by TVA as potential probes. At the 46 m (150 foot) probe location, potential readings were taken for test frequencies from 40 to 5000 Hz. The results of the measurements made while the overhead ground wires were insulated from the tower, are given in Appendix G, Tables G-12 to G-16. The last tables of this appendix (G-17 to G-19) give the ground resistances of all the 500 kV transmission towers between Johnsonville and Cumberland. These towers were measured by TVA personnel a few years after the construction of the line and towers 125 to 135 were remeasured in October 1981 during the staged fault tests. 10-32 The measurements were made using a direct-reading Megger Instrument with the return current and potential electrodes in opposite directions (along the line) at 60 to 70 m from the measured tower. As will be shown later in this section, during such measurement conditions, the measured resistance is generally lower than the real value by approximately 5% (assuming a problem-free measurement). Tower Leg Profile to Adjacent Leg Figure 10.27 500 kV Tower Leg 10.3.8 ANALYSIS OF THE HIGH FREQUENCY TESTS The measurements contained in Appendix G, Table G-16, made while the ground wires were insulated, established that there is no significant inductive component in the ground impedance of tower 130. In fact, a slight decrease in the apparent ground impedance was observed as the freguency was increased from 40 Hz to 5000 Hz. This effect is attributed to some waveform distortion in the signal of the variable frequency power unit due to the relatively high load current. The effect of connecting the ground wires to tower 130 and to 20 other adjacent towers (10 on each side of tower 130), appeared very clearly in the measurement results (see Table G-9). The apparent ground impedance, as measured from tower 130, increased regularly as the frequency was swept from 40 Hz to 20,000 Hz, indicating the presence of a significant inductive component. A simple calculation reveals that none of the resistive and inductive components of this impedance could have been constant while the frequency varied. This is to be expected if one examines Figure 10.28 which is an equivalent circuit of the electrical network which was measured. With dc or extremely low frequency currents, the equivalent combined ground impedance reduces to a pure resistance. As frequency increases, the ground wires introduce an inductance which is greatly influenced by the tower resistance values. However, when the frequency reaches a value which makes the first span impedance significantly larger than tower 130 ground resistance, the measured impedance is determined mainly by the resistance of tower 130, and becomes therefore almost a pure resistance. 10-33 a 139 Tower No. Average Span = 320 m Zg=r+jtoJl = Per Span Figure 10.28 FREQUENCY (Hz) IMPEDANCE (J]/Kra) r 0.433 0.889 1.931 1.961 1.992 2.006 1.248 2.051 2.009 1.364 2.414 1.538 4.666 2.196 2.476 7.510 9.016 3.011 1.200 0.269 1.218 0.521 60 1.238 0.767 70 1.246 100 1.274 200 400 800 1.871 1000 2.031 11.14 2000 2.792 21.54 8.289 10000 13.17 20000 20.91 30000 Table 10.6 Impedance of Ground Wires IMPEDANCE (fi/mile) r ko 50000 (10501) Equivalent Circuit with Ground Wires Connected 20 5000 UO 27.34 38.13 45.89 85-93 162.3 236.7 381.9 3.269 4.494 13.34 21.19 33.65 44.00 61.36 0.839 1.235 1.431 3-885 14.51 17.93 34.67 73.85 138.3 261.3 380.9 614.7 Ground Wire Impedance The measurement results have been plotted in Figure 10.29 and can be compared with the calculations made using the equivalent circuit of Figure 10.28 and the ground wires equivalent impedances as determined by computer program LINPA (see Appendix A). These impedance values, calculated for various frequencies and assuming a 300 fi-m average soil resistivity, are given in Table 10.6. lO-S1* 10 Computed Ground Impedance Z apparent — 129 Tower 131 132' Measured Ground Impedance 38 139 1^0 FREQUENCY (Hz) Figure 10.29 Measured and Computed Ground Impedances Good agreement exists between the measured and computed values for frequencies between 200 and 400 Hz. Outside this frequency range the agreement is in the order of ± 20% which is acceptable for this kind of measurement. It should be noted that the measured values exceed the calculated ones for frequencies less than 300 Hz and are constantly lower above that frequency. There are several factors which can explain the differences observed between the measured and computed values. Of these, it is believed that the calculated ground wires impedance value is the major factor, mainly because the impedance calculations are based on Carson's equations [20] which assume an infinite line with a uniform soil for the earth return, while the actual ground wires are finite in length (21 spans) and are grounded at intermediate points. 10.3.9 ANALYSIS OF THE LOW-FREQUENCY TESTS The low-frequency tests consisted essentially of earth potential profile measurements along parallel and orthogonal directions to the transmission line at tower 130. These potential measurements may easily be converted into fall-of-potential curves which can be used to determine the tower ground impedance (see the fall-of-potential measurement technique described in Chapter 9). This step was not necessary in this case, because the actual grounding systems were first analyzed using computer program GTOWER. It was then simple to establish, from the earth potential computations, that the measured ground impedance value at a distance of approximately 63 m (215 feet) from the tower center, is about 5% lower than the true value. Therefore, the ground impedance value of tower 130 was estimated to be about 7.1 fi when the ground wires are insulated from the tower and 1.38 fi when they are connected to tower 130 and to 20 other adjacent towers. 10-35 These ground impedance values were then used to determine the potential as a percentage of the total ground potential rise (GPR) of the tower (see Appendix G, Tables G-8 to G-16). Modelling of Tower 130 Grounding System The transmission line structures of the 500 kV line between Johnsonville and Cumberland are four-legged lattice towers with a square base. A typical TVA 500 kV tower is shown in Figure 10.30(a). The tested tower 130 is 31 m (102 ft) to the cross-arm and has a 9.3 m (30 ft) square base (Figure 10.30(b)). Each tower leg is supported by an off-center, pyramid-like, earth anchor (Figure 10.31). The earth anchor is a square base made of U-shaped steel beams resting on four triangular plates. Figure 10.31 shows the earth anchor details and dimensions. I nsu1ated Ground Wire (a) 500 kV lover Figure 10.30 (b) Tower No. 130 Base A typical TVA 500 kV Tower The four earth anchors are the only metallic elements of tower 130 which are in contact with the soil. Therefore, these earth anchors were also the transmission tower grounding system. Computer program GTOWER accepts horizontal and vertical cylindrical ground conductors only. Thus, it was necessary to develop an equivalent grounding model of the earth anchor. This is relatively easy if the following rules are observed: • The equivalent model should occupy about the same soil volume as the original system. • Grounding elements which are close to points where potential computations are desired should be replaced with modeled elements which are as similar as possible to the originals. • Closely spaced conductors should be replaced by an equivalent conductor based on the GMR technique.• • Conductor diameter has a small effect on the grounding performance. 10-36 It is possible to derive several equivalent models of any given system which will produce practically identical results. Tower Ground Line SIDE VIEW TOP VIEW Figure 10.31 300 kV Tower Leg Earth Anchor The model selected for this study is depicted in Figure 10.32. As shown later, the computation results obtained using this equivalent model were only a few percent different from those calculated using the original earth anchor (as computed by computer program MALT [15] which accepts inclined ground conductors). 0.5' — SIDE VIEW 5- 5 1 y Tower Figure 10.32 Equivalent Earth Anchor Grounding Model 10-37 Computer program GTOWER was used to analyze this equivalent grounding system when embedded in one of the ten soil models determined in Table 10.5. In addition to the tower ground resistance, two earth surface potential profiles taken along the directions shown in Figure 10.30, were calculated by GTOWER. Soil models numbered 6, 7 and 8 represent the average soil at the tower site with model 8 being the reference for the comparisons with the measured results. Ground Wires Connected to the Towers As explained in Section 10.3.8, ground impedance measured at 120 to 140. The computations with a phase angle of 16.80 at there is a significant inductive component in the apparent tower 130 when the ground wires are connected to towers made to develop Figure 10.29 yielded a ground impedance a frequency of 70 Hz (used during measurements). The potential profile measurement results are shown in Figure 10.33 in percent of the tower ground potential rise (GPR), i.e., the tower ground impedance times tower current. The measurements can be compared to the computation results from GTOWER based on soil model 8. Note however that the results from GTOWER were converted to percent taking into consideration the inductive component of the ground impedance. The procedure used to carry out this conversion is as follows. COMPUTED RESULTS (with Inductance of transmission line) ground taken into consideration Center Profile (soil model 8) Leg Profile (soil model 8) MEASURED RESULTS Center Prof Ile TVA 01 Right-Side Leg Profile TVA Left-Side Leg Profile TVA 1 Mode 1 DISTANCE FROM ORIGIN OF PROFILE (feet) Figure 10.33 Measured and Computed Potential Profiles (Ground Wires Connected to Tower) 10-38 The tower ground impedance Z was first determined from the tower ground resistance R computed by GTOWER using the relation Z = R/cos(16.8°). The converted earth surface potential values in percent of the GPR were then calculated by dividing the computed potential values by 100 ZI where I is the tower fault current. The difference between the values based on the ground impedance and those based on the ground resistance is in the order of 2 to 3% only. Figure 10.33 shows that there is very good agreement between measurements and computations for both potential profiles. SOIL MODEL CALCULATED (tt) Pi (£2-m) P 2 (fl-m) h(m) GTOWER MALT 1 250 130 1.10 6.5 - 2 311! 197 3.20 13.8 - 3 3*1 ft 113 0.8ft 5.7 U 352 162.5 2.1)7 11.5 - 5 280 120 2.56 9.0 - 6 308 150 1.76 8.35 - 7 296 163.5 1.61 8.6 - 5.4 8* 325 132 1.95 8.3 7.9 9 300 1 ^3 1.40 7.5 7-1 10 3ft 1 132 1.52 7.3 6.9 Note * Used as reference Table 10.7 Computed Ground Resistances Ground Wires Insulated From the Towers The resistivity analysis has shown that the test site soil structure was complex. That is, a multi-layer type, probably with sloping layers. It also developed three average soil models from the five base soil models constructed from the resistivity traverse measurements. It is interesting to note that despite the differences between the base soil models, the potential profile computations show essentially two types of performance as illustrated in Figure 10.34. Soil models 2, 4 and 5 gave identical to those of model 3, of 35% (with respect to the and 3 are comparable with a very similar results (model 2 results, which are practically are not shown) and are characterized by an earth potential GPR) at the center of the tower. The results from models 1 tower center potential in the order of 47%. The computed tower ground resistances however, show larger differences between the base models than those observed in the case of the potentials expressed in percent of the GPR. These resistance values are shown in Table 10.7. Note that the ground resistances based on the detailed representation of the earth anchor (use of computer program MALT), are about 5% lower than those obtained with the equivalent earth anchor model. The average difference between the potential profiles computed using both earth models is also in the order of 5%, as shown in Figure 10.36. The tower ground resistance as determined by GTOWER is practically the same (8.4 ohms) for all three average soil models (6, 7 and 8). For the reference soil model 8, which represents the average of SE5 resistivity measurements, the calculated resistance is 8.3 ohms, i.e., about 17% higher than the measured 7.1 ohms. The detailed representation of the earth anchor leads to 7.9 ohms, about 11% higher than the measured resistance. Therefore, the agreement between measured and computed ground resistances is considered satisfactory. 10-39 100 100 COMPUTED RESULTS so. Model 8 90 Mode 1 6 Model 7 cr SOIL MODEL cr a. Ijll 70 h CL 0 **» O ---------------------- n Pi o o C 60 Mode 1 1 2 3 6 o < u. CL 5 60 Pi(ft-m) 250 316 3V6 352 2So h(feet) P? Ul-m) 130 197 113 162.5 120 3.6 10.5 2.75 8.1 Leg Profi1e MEASURED RESULTS 70_ « * Center Profile TVA 07 Center Profile TVA 05 + Center Profile TVA 06 60- o 50 111 HO Model 7 Leg ProfiIe (model 8) $ c o 4^ Model 8 Model 6 80- 80 Center Profi1e Leg P rof 1 1 e (model 6) 50_ Center Profile TVA 09 O Right-side Leg Profile TVA 08 4 Left-side Leg Profile TVA 08 Leg Profi1e (model 7) LU H o a. 8.6 40“. '6.6325il3m r5- 3' | 296i1-'m 163.5n-m Model 7 cr < 30 LU cr < Hi 20 20’ - ->---- '---- 1---- 1---- 1---- 1---- 1---- 1---- 1 20 60 60 80 100 '---- 1---- ■ 120 I 1 160 T- I 160 0 DISTANCE FROM ORIGIN OF PROFILE (feet) Figure 10.34 Computed Potential Profiles (Soil Models 1 to 5) (Insulated Ground Wires) I | 20 I j 60 I J 60 i--------------- 1---------1---------------1------------<—-------T--------- ,-------------- 1--------1--------------- 1— 80 100 120 160 160 DISTANCE FROM ORIGIN OF PROFILE (feet) Figure 10.35 Measured and Computed Potential Profiles (Insulated Ground Wires) COMPUTED RESULTS (based on model 9) Center ProfiIe GTOWER Program Leg Profi1e Center Prof!le MALT Program Leg P rofi1e MEASURED RESULTS Profiles (TVA 05 to TVA 09) SOIL MODELS Profi1es based on Mode! 10. Mode 1 9 Model 10 DISTANCE FROM ORIGIN OF PROFILE (feet) Figure 10.36 Measured and Computed Potential Profiles Compatible Soil Models (Insulated Ground Wires) The measured and computed earth potential profiles are shown in Figure 10.35. Good agreement exists between test and calculated results for the leg potential profile and for the portion of the center profile which is close to the tower (1 to 10%). This agreeement decreases as the distance from the tower increases. Similar conclusions can be made regarding Figure 10.36, which shows the results of soil models 9 and 10 (labelled as "compatible models" in Table 10.5). Note that the potential values calculated which are based on models 9 and 10 are practically identical to those obtained from soil model 6'. This result confirms that several equivalent soil models can be used to suitably represent a grounding system buried in a complex soil structure. These models however, can not be constructed arbitrarily and must be based on a sound analysis of the resistivity measurements. 10-41 10.3.10 ANALYSIS OF THE TVA STAGED FAULT CURRENT MEASUREMENTS At the time of the 500 kV staged fault tests near tower 130, the equivalent positive and zero-sequence networks of the power system were as shown in Figure 10.37. Z^^I+jJS.S Z0=0.20+j0.61 TEST SITE 1o=0.03+j1.19 LEGEND ----- Equivalent Circuit ____ Actual Circuit Positive Sequence Impedance Zg~ Zero Sequence Impedance Z| > 950 NOTE: ALL IMPEDANCES ARE IN % ON TOO MVA BASE VOLTAGE = 500 kV Figure 10.37 Equivalent Positive and Zero-Sequence Network A simple calculation predicts a fault current of approximately 14,000 A for a phase-to-ground short-circuit near the test site. The fault contributions from Johnsonville and Cumberland are 4,300 A and 9,700 A respectively. However, this useful calculation does not reveal the distribution of fault current between the faulted tower and the overhead ground wires. In many cases, this can be approximated using the simplified equations of Chapter 6. These equations assume that the contribution to the fault from both sides of the faulted phase is known. If this approximation leads to significant errors, computer programs such as PATHS (see Appendix D) can be used to accurately determine the fault current distribution between towers and overhead ground wires. In all cases however, it is necessary to determine the self and mutual impedances of the transmission line conductors (faulted phase and ground wires). The analysis of the staged fault tests was carried out using the computer program PATHS developed within the scope of this EPRI project. A description of the analytical basis of this program is presented in Chapter 6. Construction of the Equivalent Transmission Line Circuit In order to analyze a faulted transmission tower using the computer program PATHS, it is necessary to represent each of the faulted transmission line terminals by an equivalent single power source or load. The transmission line may be single or double-circuit and may 10-42 carry a load current prior to the fault. The load current and its direction of flow are described by the magnitudes and angles of the vectors representing the driving source voltages at each terminal. The equivalent circuit of the TVA 500 kV power network used for the staged fault tests is shown in Figure 10.38. The impedances of the transmission line equivalent terminals were determined from the data given in Figure 10.37. The positive, negative and zero-sequence impedances were converted into phase quantities using the relation Zphase = (1/3)(Z1+Z2+Z0). The transmission line conductor impedances were calculated by the computer program LINPA described in Appendix A. The line parameter calculations were made assuming a 300 f2-m average soil resistivity and the line geometry shown in Figure 10.39. This figure also gives the computed line impedances for the equivalent 4-wire transmission line (obtained as a result of phase bundle and ground wire reductions into equivalent single conductors). The terminal ground resistances were selected as 0.1 in accordance with the test results which show that the Johns onville and Cumberland station ground impedances are very low (below 0.1 Q). These resistances have negligible effects on the calculations of this study. Figure 10.38 Equivalent Faulted Transmission Line Circuit The analyses were initially carried out assuming that the source voltage at each end of the transmission line was 500 kV (phase to phase). In fact, the transmission line was carrying about 480 A from Cumberland to Johnsonville prior to the staged fault tests. Thus, other PATHS runs were made, this time with the magnitude of the driving voltage at the terminals equal to the measured values during the tests. The computed net current in the phase conductors was 500 A, flowing from Cumberland to Johnsonville. Only small differences were observed between the computed results obtained based on the equal and unequal voltage assumptions, as shown by the values given in Table 10.8 The computations were made for the bonded ground wire condition, i.e., with the ground wires solidly connected to tower 130 and to 20 other adjacent towers. The condition with the ground wires insulated from the towers except at the towers where the ground wire insulators flashed over was not analyzed because the flashovers occured randomly on each side of tower 130. Consequently, it was not possible to determine from the test results 10-43 which towers were conducting current into ground. However, the test results indicated that several towers outside the zone which was monitored (i.e., the zone defined by towers 125 to 135) did flash over. It is believed that the random nature of the location of the ground wire insulator flashover is due to the wide variations in dielectric strength of the insulator assembly. This strength is estimated to be 10 to 20 kV depending on the air gap distance provided by the two clipped washers of the insulator [13]. 10.1m 10.1m Ground W1re 7 No.9 A1umowe1d Phase Conductors 954 kC mi] (1.5') 12.2 m 45/7 Strands ACSR (RAIL) B1 @ i ® B2 Phase B Phase A Phase C MATRIX OF CONDUCTOR IMPEDANCES (fl/mi le) A (A1 + A2 + A3) B (B1 + B2 + B3) C (Cl + C2 + C3) A 0.131 + jl.136 B 0.093 + j0.584 0.129 + j1.136 c 0.091 + jO.501 0.093 + jO.584 0.131 + j1.136 D 0.092 + jO.558 0.092 + j 0.570 0.092 + jO.558 Figure 10.39 0 (G1 + 02) 1.992 + j1.235 Transmission Line Conductor Impedances The measurement results and the related computation results are summarized in Table 10.8 for the bonded ground wire condition, and in Table 10.9 for the insulated ground wire condition. The measurement results were determined from the oscillograms which were recorded at both station sites and at the test site during the four staged fault tests carried out during the last week of the month of October 1981. These oscillograms are shown in Figure 10.40, 10.41 and 10.42. Figure 10.40 shows the oscillograms recorded at the test site near tower 130. As mentioned, no meaningful oscillograph traces were obtained during test No. 1. Tests No. 2 and No. 3 were made when the overhead ground wires were solidly connected to the 20 towers adjacent to tower 130 and during almost identical power system conditions. Therefore, one would expect similar values of current in the same tower for each test. This is not the case, as shown in Table 10.8. This table shows considerable difference between the measured and calculated results at towers 134, 135, 133 and 131. Since the fiber optic cables used to trasnmit the measurement signals from these towers to the recording site were repaired after being cut and because good agreement exists between measured and calculated results at all locations which did not experience fiber optic cable damage, it is concluded that the measured currents at towers 132 to 135 for tests No. 3 and 4 are inaccurate, due to equipment failure or calibration error. 10- 44 Table 10.8 gives a summary of the measured and calculated results assuming: a- Equal source voltage magnitudes at Johnsonville and Cumberland (500 kV phase to phase) and ignoring the 11.5 mH coils in series with the ground wires at these stations. b- Equal source voltage magnitudes with the 11.5 mH coils in place. c- Unequal source voltage magnitudes at the terminals with the 11.5 mH coils in place. TEST 1 TEST 2 RES U L T S COM P U T E D MEASURED RESULTS TEST THREE PHASE CIRCUIT 3 a** I 135 (A) - - - i n1* (A) - 375 112* 301 I 133 (A) - - 225* 539 I 132 (A) - - 300* I 131 (A) 750 T 129 (A) 128 423 b** 412 c** SINGLE PHASE CIRCUIT a** 435 315 - 569 - - b* 415 299 539 629 c** 438 316 571 665 893 627 297 537 627 664 - 750 840 842 892 - 843 - - 913 915 969 - 916 969 (A) 825 - 1110 1113 1179 - 1114 1179 1 127 (A) 450 - 628 - 630 667 (A) - - 800 849 - 802 849 T 125 (A) - - 712 630 802 711 667 I 126 753 - 711 753 12410 12421 13150 - 12407 13139 10960 10973 11616 - 10959 11605 10274 10304 10968 10307 10916 I TTotal Fault (a; 12375 12300 I PHGW (a: - f V Orthoqonal tv; - * 10800 9125 t V Parallei (v; - * 10000 i no (Ai 10800 1500 2380 1447 1451 1536 - 1452 1537 2400 2176 1973 2188 - 1893 2097 9264 8930 9240 9327 9340 10865 - 9329 10806 9000 8880 8880 9307 9320 10375 - 9309 10306 (A 6048 6050 6240 7920 7755 8216 7762 8595 (A 85 95 90 V Phase A“N (kV 308 - IReturn 0HGW (a; 2A80 I Residua1;31 □ (A 1 Phase A (A' IN 500 kV IN 161 kV IReturn OHGVj (A - 1575 308 308 288 288 319 - 288 319 1100 900 725 690 731 - 705 643 IResIdual 3Io(A 3180 3200 3120 3085 3082 2278 - 3079 2335 I Phase A (A 3720 3720 3600 3105 3102 2778 - 3099 2835 IN 500 kV (A) 2200 2300 2300 2603 2601 2747 2628 2402 IN 161 kV (A) 288 288 264 288 264 V Phase A-N (kV) - 264 264 1200 264 - Notes ■' Inaccurate (or unusable) test results ** Assumption a = equal terminal voltage /no 11.5 coils Assumption b = equal terminal voltage/and 11.5 mH coils Assumption c * unequal terminal voltages/and 11.5 mH coils f This is the GPR of tower No. 130. Table 10.8 Banded Ground Wires Test Results and Computations The above three conditions were analyzed using computer program PATHS based on the actual three-phase circuit and assuming a single-phase circuit (phases B and C ignored). Minor modifications to the input data routines of PATHS were necessary to run some of these cases. Detailed examination of Table 10.8 shows good agreement between measured and calculated results. From an engineering point of view, computations based on the three assumptions, a, b, and c are equivalent. The computed values of current returning from earth to the transformer neutrals at both stations are consistantly higher than the measured values while the reverse is true for the current returning through the ground wires. Therefore, it is suspected that the mutual impedances between the phase conductors and the ground wires are slightly higher than those used in the computations. 10-45 Particularly good agreement was obtained between measured and calculated ground potential rise (GPR) of the faulted tower which is the reference value in the transmission line grounding analysis. The measured potential rise values (shown in Table 10.8) were obtained with respect to a reference point some 60 m (200 ft) away from tower 130. As already discussed in this chapter, this situation leads to a measured GPR approximately 5% lower than the true value. Therefore, the comparison should be made using a GPR 1.05 times the measured value. MEASURED RESULTS TEST 4 t t Notes Table 10.9 I 135 (A) I 13^ (A) I 133 (A) I 132 (A) - I 131 (A) - I 129 (A) - I 128 (A) - I 127 (A) I 126 (A) T 126 (A) I Total Fault (A) 13100 T 0HGW (A) 11000 V Orthoaonal (V) 13125 V Parallel I 130 (V) (A) 13750 2100 I Return 0HGW (A) 2360 I Residual ; 3Io (A) 9240 I Phase A (A) 8928 IN 500 kV (A) 5856 IN 161 kV (A) V(J) A-N (kV) I Return 0HGW (A) 785 I Residual ; 31o (A) 3240 I Phase A (A) 3680 IN 500 kV (A) 2250 IN 161 kV (A) 1200 V4> A-N (kV) 375* - 750* - 488** _ 90 308 264 * Inaccurate test results ** In previous tests no osciHogram traces were obtained for tower 126. Therefore, this value is questionable. This is the GPR of tower No. 130. Insulated Ground Wires Test Results The measured results of test No. 4 corresponding to the insulated ground wire case, are summarized in Table 10.9. Because the towers which conducted fault current to earth were not identified, it was not possible to simulate the test using program GTOWER. However, the measurement results indicate that the major difference between this and other tests is in the increase in the proportion of fault current flowing through tower 130 (about 2000 A compared to 1500 A in the other tests). Computer simulations show that as the number of towers available to conduct current to earth decreases on each side of the faulted tower, the magnitude of the faulted tower current increases (assuming that the equivalent source impedance is several times higher than the tower impedance). This suggests that the number of towers which conducted fault current into the earth during test No. 4 was less than 10 out of 20 towers. 10- 46 Consequently, it is necessary to apply safety factors to the computed current (assuming continuously bonded ground wires) of a tower in a transmission line equipped with insulated ground wires. Based on the above data, a safety factor of 1.5 is suggested. An alternate, and perhaps more rigorous, method consists of computing the tower fault currents initially assuming that flashovers only occurred at the two towers immediately adjacent to the faulted tower. The calculations are then repeated, each time adding a new pair of towers, until the computed voltage accross the ground wire insulation of the closest noncurrent-carrying neighbouring towers becomes less than the minimum dielectric strength of the ground wire insulators. TOWER 135 CURRENT 1 p.u. = 2400 A SPAN GW CURRENT (135) 1 p.u. = k80Q A TOWER 134 CURRENT 1 p.u. = 2400 A TOWER 133 CURRENT 1 p.u. = 2400 A TOWER 132 CURRENT 1 p.u. = 2400 A TOWER 131 CURRENT 1 p.u. = 2400 A TOTAL TOTAL FAULT (lt) 1 p.u. = 2400 A 0HGW CURRENT IN 0HGW 1 p.u. = 14400 A TOWER 128 CURRENT 1 p.u. = 2400 A TOWER 127 CURRENT 1 p.u. = 2400 A TOWER 126 CURRENT Ortho. TOWER 130 GPR 1 p.u. = 40 kV TOWER 130 GPR Paral. 1 p.u. = 40 kV SCALE: 1 p.u. Figure 10.40 TEST No. 2 TEST No. 3 TEST No.4 (Ground Wires Current and Voltage Oscillograms Measured at Tower Site 10- h7 Insulated) TOWER 130 SITE OHGW p.u. = 800 A Residual 1 p.u. = 2A00 A 500 kV Neutra 1 p.u.= 2400 A 161 kV Neutral p.u. = 320 A Phase A 1 p.u. = 2400 A Phase A-Neutral 1 p.u. = 176 kV 1 p. u. 1 p. u. TEST No. 1 TEST No. 2 SCALE Figure 10.41 TEST No. 3 TEST No. 4 {Ground Wires Insulated) Current and Voltage Oscillograms Measured at Cumberland Station 10- 48 OHGW Residua 1920 A 500 kV Neutral 1 p.u. = 960 A 161 kV Neutral NOT RECORDED NOT RECORDED 1 p.u. = 960 A Phase A 1 p.u. 1200 A Phase A-Neutral p.u.= 132 kV TEST No. p.u. 1 p.u. SCALE Figure 10.42 TEST No. 2 TEST No. 3 TEST No. k (Ground Wires Insulated) Current and Voltage Oscillograms Measured at Johnsonville Station 10- bS 10.4 THE OH/AEP 500/765 kV TRANSMISSION LINE MEASUREMENTS I In April and May 1979, the American Electrical Power (AEP) conducted a series of staged fault tests on the Kammer-Marysville 765 kV line. A total of four towers were faulted. At each of the four towers, two guyed-vee and two lattice structures, measurements were made of fault currents, earth components of fault currents and earth potential profiles. Ontario Hydro (OH) participated with AEP in the earth potential profile measurements. Later, a series of tests was also performed on an isolated 500 kV lattice tower adjacent to Ontario Hydro's outdoor high voltage test facility at Kleinburg, Ontario. These measurements have been described and discussed in detail in a recent comprehensive paper [7]. Only the key aspects of the measurements, including a comparison of the test results with analytical computations, are presented in this section. An interesting feature of the measurements is a study of the effectiveness of a special ground potential control ring (GPC) installed as part of the tower grounding, to reduce step and touch voltages. 10.4.1 THE AEP MEASUREMENTS Figure 10.43 shows the AEP Kammer-Marysville 765 kV transmission line and the locations of the four test sites. Two single phase to ground faults were staged at each tower while the Kammer station was open-ended during the tests. The fault current fed from Marysville station during each fault is given in Figure 10.43. Lattice Towers Guyed-Vee Towers > 4860 A 6330 A 2645 A 2650 A u~\ SO 4920 A 6424 A 2645 A 2400 A _l Tower > I 515 T t ^0 I LU _J 486 36 1 13 s 24 Km : ' (15 mi}': , 56 Km (22 mi ) 230 Km “(143 mi ) 237 Km (147 mi 244 Km (151 Figure 10.43 mi) The AEP 765 kV Test Sites (Redrawn from [7]) Figure 10.44 shows a plan view of the guyed-vee towers and the cross section of the GPC ring installation around guy anchor D of tower 13. This figure also shows the directions of the earth potential profiles B, D and E measured at tower 13. Only the D potential profile was measured for tower 36 which was not equipped with the GPC ring. 10-50 ORIGIN OF PROFILE Profile B ^ Profile E Kammer Marysvi11e 5-5 m ORIGIN OF PROFILE Tower Base TOP VIEW Tower Guy Anchor Wi re Profile D SIDE VIEW Connecting Spoke 1.8m Diameter Guy Anchor Diameter No. k Copper Ring Figure 10.44 Plan View of Guyed-Vee Towers (Redrawn from [7]) Several resistivity measurements were taken at various locations near the towers prior to and following the tests. Significant variations in the measured values along various traverses were observed. These variations are probably a result of the hilly terrain at the towers. Therefore, the average apparent soil resistivities of Figure 10.45 were used to determine an equivalent two-layer soil model at each tower site. Computer program RESAP [16] was used to perform the resistivity computations. RESAP is the more general SES computer program from which RESIST was derived. The computed characteristics of the equivalent soil models, also given in Figure 10.45 were used by the computer program MALT [15] to calculate the tower resistance and the earth potential profiles. MALT is a general SES grounding program used as a reference to develop the program GTOWER. The results of the measurements and MALT computations are shown in Table 10.10 and in Figures 10.46 and 10.47. Because the tower grounding system consists essentially of horizontal and vertical elements, the computation results of MALT and GTOWER are practically identical. 10-51 Pi = 60 Q-m P2 = 250 Q-m h = 3 ni Towe r 1 3 ____ Tower 36 PROBE SPACING Retire 10.45 (m) Average Apparent Soil Resistivities Around Guyed-Vee Towers (Redrawn from [7]) TOWER No. RESISTANCE (fi) Measured Computed 13 2.3 2.34 36 2.0 2.08 Table 10.10 Guyed-Vee Tower Resistances Very good agreement between computed and measured resistances was obtained for both towers. Good agreement between the computed and measured potential profiles is evident except in profile B of tower 13. The differences between measured and computed values are in the order of 10% and are probably a result of the soil structure at the profile B location which was significantly different from the average soil structure assumed in the computations. The four staged fault tests at towers 13 and 36 indicated that only 20 to 30% of the total fault current entered the tower ground. Most of the fault current flowed in the ground wires which were connected to the faulted tower. The measurements conducted at the four-legged lattice towers 486 and 515 lead to similar observations as at the guyed-vee towers. However, the agreement between computed and measured values, rated between good and acceptable for most results, is poor for 2 of the 8 potential profiles measured (discrepancy of 20 to 50%) and for the tower resistance values when the GPC ring is not connected to the tower. 10-52 Measurements C£ o a. Measurements .MALT Computations L. <D § Profile D o d-P Profile E < O Q. Profile B X Iq; < UJ LATERAL DISTANCE (m) Figure 10.46 100 . Earth Potential Profiles for Tower 13 (Redrawn from [7]) MALT Computations Measurements Profile D LATERAL DISTANCE Figure 10.47 (m Earth Potential Profile for Tower 36 (Redrawn from [7]) In all cases however, the measurements and computations show that when the GPC ring (see Figure 10.48) is connected to the latttice tower, the reduction in touch voltage (compared to the unconnected case) is substantial. 10-53 GPC Ring (actual Tower Foot 1ng No. GPC 4 Copper Ring omputer Mode 1} GPC Ring TOP VIEW — Ground Rod Potential Profile B Potentia 1 Profile C \ Potentia Potent I a 1 Profile A SIDE VIEW Figure 10.48 Plan View of 765 kV Lattice Tower Footing (Redrawn from [7]) 10.4.2 THE OH MEASUREMENTS The AEP measurements clearly indicate that to achieve good agreement between measured and computed values, the soil structure at the test site must be accurately determined and the measurements must be performed under controlled conditions insensitive to or isolated from interference from energized transmission line conductors and other sources of electrical noise. This objective was met by performing a series of tests on an isolated 500 kV lattice tower situated approximately 300 m from Ontario Hydro's outdoor high-voltage test facility at Kleinburg. The legs of this tower are insulated from the tower grounding and the concrete reinforcing bars by insulating plates, as shown in Figure 10.49. The GPC ring of No. 4 copperweid wire and the ground rods were bonded to the steel. INSULATOR WINDOW TYPE CT Figure 10.49 500 kV Tower Leg at Kleinburg (Courtesy of Ontario Hydro) 10- 5k The GPC ring and the four ground rods were electrically connected to the steel at each leg by removable bonding straps. This grounding scheme, illustrated in Figure 10.50, permitted measurement of the tower resistance and earth surface potentials for various grounding arrangements. Thus, it was possible to examine the contributions from the concrete pier rebars, ground rods and GPC ring either alone or in combination. Earth surface potential measurements were made using earth contact pads weighted down by 15 kg (33 lb) concrete cylinders as shown in Figure 10.51. GPC Ring (Actual) Potential Prof!le TOP VIEW GPC Ring (Computer Model Tower Footing ------ 3m Ground Rod Connecting St rap SIDE VIEW Figure 10.50 Plan and Cross Sectional Views of the 500 kV Tower Footing (Redrawn from [7]) Figure 10.51 Earth Surface Potential Measurement Pads (Courtesy of Ontario Hydro) 10-55 Core drillings in the vicinity of the test tower revealed that the soil is clay to 30 m (98 ft) depth. Computer analysis of the apparent resistivity data, shown in Figure 10.52, yielded a two-layer soil model consisting of a 30 Q-m resistivity top layer, 15 m (49 ft) deep, and a 100 . fi-m resistivity bottom layer. This two-layer soil model was used to conduct the analyses with the computer program MALT. 100 . Computations (RESAP) Measurements PROBE SPACING (m) Figure 10.52 Resistivity Measurements at Kleinburg Test Site (Redrawn from [7]) The measurement and computation results for the six different grounding arrangements are summarized in Table 10.11 (tower ground resistances), and in Figures 10.53 to 10.55 (normalized earth surface potentials along the profile shown in Figure 10.50). GROUNDING ARRANGEMENT RODS ONLY Measured Resistance (A) 2.77 1.06 1.41 0.966 1.33 0.97 Computed Resistance (A) 2.78 1.07 1.61 1.03 1.56 1.00 Table 10.11 RODS AND FOOTING REBAR FOOTING REBAR RODS AND RODS, REBAR GPC RING CAGES ONLY AND GPC RING FOOT 1NG REBAR AND RING Kleinburg Tower Ground Resistance 10 -56 Ground Rod Location Ring Location Rods Only Rods + GPC Ring Measurements STANCE FROM ORIGIN OF PROFILE Figure 10.53 (m) Earth Surface Potential Profiles for Ground Rods and GPC Ring (Redrawn from [7]) Ground Location Location ___ Rebar Only ------ Rebar + GPC Ring . Measurements DISTANCE FROM ORIGIN OF PROFILE (m) Figure 10.54 Earth Surface Potential Profiles for Rebar Cages and GPC Ring (Redrawn from [7]) The computations take into account the entire grounding configuration as illustrated in Figure 10.50. In computing the tower grounding performance for the ground rods alone, the rebar cages and the GPC ring are consi dered as in situ metallic elements not bonded electrically to the ground rods. This was necessary because these elements, especially the rebar cages, act as a current path to the ground rods and noticeably influence the measured results (see the discussions and conclusion at the end of [7]). 10-57 Although not shown in the figures, the potential profiles were measured and computed over a lateral distance of 24 m (79 ft). Agreement between measurements and computations is remarkably good over the entire profile. Similar excellent agreement also exists between measured and computed ground resistances of all grounding configurations. Ground cc CL. Location o 4- O <3^ Location ---------• Rods + Rebar Rods + Rebar + GPC Ring Measurements < O Q_ < UJ DISTANCE FROM ORIGIN OF PROFILE (m) Figure 10.55 Earth Surface Potential Profiles for Ground Rods With Rebar Cages and GPC Ring (Redrawn from [7]) The success of this measurement/computation program is due to: • Accurate measurement of the apparent soil resistivity at the test site and derivation of a suitable equivalent two-layer soil model. • Absence of electrical interference and noise. • Good test procedures and appropriate measuring equipment. • Detailed knowledge and representation of the metallic elements forming the tower grounding system. In bonding the GPC ring to the tower, the touch voltage for a person standing 1 m (3 ft) away from the tower leg decreases from approximately 60% of the tower potential rise in the "rods only" case to about 20%. However, since the tower potential rise, assuming constant current, also decreases when the GPC ring is bonded to the tower (because the ground resistance has decreased), the actual reduction of the touch voltage is even greater. The authors of the OH-AEP test program also conducted a sensitivity analysis using the computer program MALT, to determine the influence of the fundamental parameters on the grounding performance of a 500 kV lattice tower equipped with one or two GPC rings. Their main conlclusion was that it is possible to reduce the touch voltages to less than 5% percent of the value which exists in the absence of the GPC rings. 10- 58 10.5 THE EDF ROCXET-TRIGGEERED LIGHTNING MEASUREMENTS Electricite de France (EDF) has recently completed an extensive lightning measurement program. This program started in 1973 and was still active in 1981 [8,9,10]. During this eight year program at the Saint Privat d'Allier test site in France, 94 lightning flashes were triggered by means of the rocket-wire method described by Newman [17]. The program included a wide variety of measurements intended to gain further understanding of the lightning process and its effect on the environment, including power transmission lines. In particular, a series of measurements were conducted to determine the response of a transmission line tower and its grounding system to direct lightning strokes. The principal facilities at the experimental station are a transmission line tower supporting a rocket-launching platform at the top (see Figure 10.56) and a ground level launching area (see Figure 10.57). Figure 10.56 Tower Structure at St. Privat d'Allier (Courtesy of Electricite de France) 10-59 I Figure 10.57 Ground Level Rocket-Launching Area (Courtesy of Electricite de France) The tower structure is 26 m (85 ft) high and is mounted on insulators (see Figure 10.58). This tower is equipped with appropriate measuring equipment to monitor the potential rise of the structure and the grounding system potential rise during direct lightning strokes. During 1977 and 1979, it was possible to measure the lightning performance of three different tower grounding electrode configurations (see Figure 10.59): • A 1 m (3 ft) radius hemisphere • A 1 m x 1.5 (3 ft x 4.5 ft) double loop electrode. • A 6 m (20 ft) long vertical ground rod. A surge generator was used at another test location, to measure the surge characteristics of three additional electrode configurations, in order to supplement the rocket triggered lightning measurements [9] (see Figure 10.59). The measurement results (current and voltage magnitudes as a function of time) were translated into the frequency domain using a Fast Fourier Transform algorithm adapted to lightning surges. This approach has the advantage of showing the response of the tower and its grounding system as a function of frequency and can be used to predict the lightning performance of the tower for any current waveform once the current is broken down into its harmonic components. < 10 -60 Figure 10.58 Tower Base and Insulator Supports (Courtesy of Electricite de France) HORIZONTAL CONDUCTOR GROUND ROD HEMISPHERE Xi F f; "7 J F 1m X (3.3') ,1 w 0.6m (2') i 15m (50') TWO-BRANCH STAR Figure 10.59 !l1l|] TTFTT IfP —T" 1 .Sm X ----------------^a. ... . ...=3: (5') y 11m * nn / (3.3') DOUBLE-LOOP GRID (Depth = 0.4m) Electrode Configuration Subjected to Surge Tests (Redrawn from [10]) 10 -61 10.5.1 EQUIVALENT CIRCUIT OF THE TOWER STRUCTURE At low frequencies, the impedance of a tower structure can easily be neglected with respect to its footing ground resistance. At very high frequencies or for surge and lightning currents, the structure may represent a significant portion of the overall impedance. There is extensive literature on the lightning response of transmission line towers. Unfortunately, no consensus has been reached concerning the optimum representation of a tower structure. There are two basic concepts; the first one, often used to establish the rate of failure of a transmission line, considers the tower structure as a surge impedance; the second approach lumps the tower structure into a localized self-inductance. The EDF lightning measurements support the localized self-inductance concept as can be seen from Figure 10.60, which gives the tower structure impedance as a function of frequency. This figure shows that the tower structure inductance is in the order of 3.5 mH, a value about half the theoretical value calculated using the formula [18]: L = ^ (10-3) where L is the inductance (mH) h is the tower height (m) r is the equivalent radius of the tower base (m) for h = 26 m, r = 2 m, the tower inductance L is 7.6 mH. Measured (26 m Tower) FREQUENCY (Hz) Figure 10.60 Tower Structure Impedance (Redrawn from [10]) The EDF measurements were obtained during actual lightning strokes and were measured with sophisticated equipment and great care. Thus, solid evidence in favor of the localized inductance concept has been obtained. However, more measurements must be conducted for other tower structure configurations and heights, in order to eliminate any doubt concerning the validity of this concept. 10-62 10.5.2 EQUIVALENT CIRCUIT OF THE GROUNDING ELECTRODES No soil ionization phenomena (soil breakdown due to excessive potential gradients) were recorded during rocket-triggered lightning tests at the tower site for current peaks in the order of 20 kA. Since the tower structure is not equipped with ground wires as in most transmission lines and is therefore isolated, it can be concluded that the ionization effect can be ignored in most cases involving high voltage lines and lightning currents in the order of 20 kA crest or less. Horizonta1 Conductor Double Loop Hemisphere FREQUENCY (Hz) Figure 10.61 Ground Impedance of Hemispherical, Double Loop and Horizontal Electrodes (Redrawn from [10]) Ground Rod Two-Branch Star FREQUENCY (Hz) Figure 10.62 Ground Impedance of a Ground Rod and a Two Branch Star (Redrawn from [10]) 10-63 The measurement results, shown in Figures 10.61 and 10.62, indicate that except for the hemispherical electrode, the grounding electrodes exhibit an inductive component which become noticeable at very high frequencies, generally above 100 kHz. The EDF results also show that the impedance of the ground electrodes is a rising function of frequency and there is no indication that an asymptotic value is reached as the surge impedance concept suggests. From the ground rod measurement results it is concluded that concentrated electrodes can be represented by a resistance in series with an inductance. This representation is valid for electrode dimensions of less than 5 to 8 m (16 to 26 ft). For a 6 m (20 ft) ground rod the measured inductance was in the order of 5 txH. For ground conductors 10 m (30 ft) and more, in length, the ground impedance increases as the square root of the frequency at high frequencies. This indicates that the ground electrode can be modelled by a line with distributed constants [10]. The ground impedance can be approximated from: Z = /R(R + jZTrfl) (i°-4) where Z is the ground impedance (ohms). R is the low frequency resistance of the electrode (ohms). L is the self inductance of the electrode (Henrys). f is the frequency (Hertz). The measurements indicated that the self inductance of a linear ground electrode conductor is in the order of 1.2 mH per meter. REFERENCES 1 - F. Dawalibi, D. Mukhedkar, "Optimum Design of Substation Grounding in Two-Layer Earth Structure - Part I, Analytical Study". IEEE Transactions, Vol. PAS-94, No. 2, March/April 1975, pp. 252-261. 2 - F. Dawalibi, D. Mukhedkar, "Transferred Earth Potentials in Power Systems", IEEE Transactions, Vol. PAS-97, No. 1, January/February 1978, pp. 90-101. 3 - F. Dawalibi, D. Mukhedkar, D. Bensted, "Measured and Computed Current Densities in Buried Ground Conductors". IEEE Transactions, Vol. PAS-100, No. 8, August 1981, pp. 4083-4092. 4 - Michael A. Sargent, Mat Darveniza, "Tower Surge Impedance", IEEE Transactions on Power Apparatus and Systems, Vol. PAS-88, No. 5, pp. 680-687. 10-64 5 - F. A. Fisher, J. G. Anderson, J. H. Hagenguth, "Determination of Lightning Response of Transmission Lines by Means of Geometrical Models", AIEE Transactions on PAS, February 1960, pp. 1725-1735. 6 - P. C. Buchon, W. A. Chisholm, "Surge Impedance of HV and EHV Transmission Towers", Report No. 78-71, Canadian Electrical Association. 7 - E. A. Cherney, K. G. Ringler, N. Kolcio, G. K. Bell, "Step and Touch Potentials at Faulted Transmission Towers", IEEE Transactions on PAS, July 1981, Vol. PAS-100, No. 7, pp. 3312-3321. 8 - R. Fieux, P. Portal, "La Station Experimentale de St. Privat d'Allier Resultats Generaux, Caractbristiques des Coups de foudre Declenchbs". E. D. F. Bulletin de la Direction des Etudes et Recherches - serie B, Reseaux Electriques, Matiriels Electriques, No. 4, 1979, pp. 39-63. 9 - R. Fieux, P. Kouteynikoff, F. Villefranque, "Measurement of the Impulse Response of Grounding to Lightning Currents", Electricite de France, Proc. of the 15th Europeen Conf. on Lightning Protection, Vol. 2, pp. K440-K450, Uppsala, Sweden. 10- P. Kouteynikoff, "Reponse Impulsionnelle des Prises de Terre aux Courants de Foudre" CIGRE paper 121-03, 1981. 11- A. J. Pesonen, "Effects of Shields Wires on the Potential Rise of HV Stations" Offprint from SAKHO - Electricity in Finland, 1980, No. 10, pp. 305-308. 12- Janos Endrenyi, "Analysis of Transmission Tower Potentials During Ground Faults", IEEE Transactions on PAS, Vol. PAS-86, No. 10, October 1967. 13- A. C. Pfitzer, G. M. Wilhoite, "Tennessee Valley Authority's 500 kV System - Transmission Line Design", IEEE Transactions on PAS, Vol. PAS-85, No. 1, January 1966, pp. 28-35. 14- G. V. Keller, F. C. Frischknecht, "Electrical Methods in Geophysical Prospecting" Pergamon Press 1977, (Book). 15- F. Dawalibi, "User Manual for Program MALT", Safe Engineering Services Ltd., Montreal, 1979. 16- F. Dawalibi, "User Manual for Computer Program RESAP", Safe Engineering Services Ltd., Montreal, 1979. 17- M. N. Newman, "Use of Triggered Lightning to Study the Discharge Channel" Problems of Atmospheric and Space Electricity, Elsevier, New York, 1965. 18- R. Fieux, "RQle des cables de Garde sur les lignes 63/90 kV", Note HM/72-4404, 1980, Electricite de France. 19- R. H. Golde, "Lightning", Vol. 2, Academic Press 1977 (Book edited by R. H. Golde). 20- John R. Cars on," Wave Propagation in Overhead Wires With Ground Return", Bell System Technical Journal, October 1926. 10-65 I CHAPTER 11 SURVEY OP NORTH AMERICAN TRANSMISSION LINE GROUNDING PRACTICES 11.1 OBJECTIVES OF SURVEY As a very important part of a comprehensive examination of transmission line grounding methods, a survey was undertaken to determine the practices and attitudes of North American power utility companies regarding the grounding of high voltage transmission lines. The questionnaire was designed to investigate specific quantitative standards used in the design of transmission line grounding systems as well as to examine the general attitudes and concerns of power utilities on the subject. These inputs were solicited to assure that the results of the work being carried out would adequately reflect the concerns, objectives and experience of the utilities for whom the work is being undertaken through EPRI. The questionnaire addressed problems related to lightning and 60 Hz fault conditions and examined safety and operating reliability considerations in the design of grounding systems. Questions were asked which would permit grounding system design practices to be related to system voltages, keraunic levels, failure rates, and perceived hazard levels associated with ground faults. Specific details regarding measurement, construction and maintenance procedures were surveyed with the objective of providing useful feedback to utilities on how their practices compare with others and to guide in the adoption of recommended standards in these areas. 11.2 SURVEY DISTRIBUTION The survey questionnaire was distributed to a total of 162 power utilities in the United States and Canada. The utilities were selected from the Electrical World Directory of Electric Utilities and from lists of utilities provided by EPRI. Only those utilities identified as having high voltage transmission lines rated 115 kV or higher were selected for participation in the survey. A total of 88 completed or partially completed questionnaires were received. In some cases, one response was identified as representing several companies under common ownership. Thus, the results represent the input of somewhat more than 88 utilities. The distribution of utilities responding to the questionnaire according to total number of employees is shown in Figure 11.1. It can be seen that the greatest portion of respondents (25%) were utilities with between one and two thousand employees and that only 6% of the responding utilities had fewer than one thousand employees. 11-1 30 . 30 25 - 25 <y> 20 15 o O' 10 5 0 FROM: TO: 0 0.5 0.5 1 1 2 2 3 3 5 5 7 7 10 10 15 15 30 TOTAL NUMBER OF EMPLOYEES (IN THOUSANDS) Figure 11.1 Size of Utilities which Answered the Questionnaire In most cases, the questionnaire was sent to the manager of Transmission Engineering or the equivalent position within the utility. A list of participating organizations is contained in Appendix E. 11.3 RESPONSE TO THE SURVEY As indicated, a total of 88 questionnaires, representing approximately 95 utilities were received. In many cases, the questionnaire was only partially completed, presumably because the respondent did not understand the question or had no involvement with the specific topic being questioned. Additional comments were provided on at least one subject by 77 utilities, and 36 utilities provided reference documentation such as internal standards, reports, design guides, or drawings. Occasionally, the response to a particular question by a respondent was ambiguous or in obvious conflict with the response to another question. In these cases, an attempt was made to resolve the ambiguity by analyzing the data further or seeking clarification from other sources. The comments received provide an excellent source of background information on most topics covered in the survey, and several additional topics or points of view not specifically addressed in the questionnaire. 11.4 QUESTIONNAIRE FORMAT The questionnaire was divided into six sections, covering various aspects of transmission line grounding practices. Section I covered general aspects including design standards, performance objectives, costs and perceptions of the relative importance of grounding. The second section dealt with pre-cons true ti on measurements and was intended to determine the extent to which utilities carry out field measurements related to transmission line grounding and the techniques used in such measurements. 11-2 I Section III dealt with engineering design aspects and questioned the extent of calculations performed relating to grounding variables used in the design equations, basic hypotheses, and the degree of satisfaction with calculation techniques. This section was subdivided to consider both lightning and 60 Hz fault conditions. Section IV addressed safety considerations and the extent to which calculations and/or measurements are carried out to evaluate the safety of grounding systems. The variables used in design calculations for safety were surveyed, as were the attitudes of utilities regarding hazard levels associated with transmission line grounding systems. Section V covered construction, measurements and maintenance and surveyed specific details of ground electrode construction materials and techniques, and subsequent testing and maintenance of installed grounding systems. The final section dealt with current practice and design standards of utilities for transmission line grounding systems. Skywire usage, structure ground impedances and the important question of failure rates due to lightning were surveyed. Most questions were of the multiple choice or single word (or number) answer type. Comments and reference documentation were requested in most sections. 11.5 ANALYSIS OF THE RESPONSES The responses to the questionnaire were initially tabulated by question number and the percentage of respondents in each reply category determined. Since most questions were not answered by all respondents, each percentage was determined as a percent of those answering the question, as well as a percentage of total respondents. In order to gain insight into the significance of grounding practices and transmission line performance, an attempt was made to correlate transmission line failure rate data with various design practices and standards. This is discussed in Section 11.5.6. The important results of the six sections of the questionnaire are summarized below. 11.5.1 GENERAL (SECTION I OF SURVEY QUESTIONNAIRE) From the responses to the questions in this section, it is apparent that power system performance under the influence of lightning disturbances is the overiding concern of utilities when transmission line grounding systems are designed. Most utilities (94%) considered grounding to be important and 62% of those considered it to be of fundamental importance with respect to the overall performance and reliability of transmission systems. Grounding represents less than 5% of the design and construction budgets of most utilities although 13% reported that construction costs directly related to grounding accounted for 5 to 10% of the total construction budget. Only one utility reported engineering costs related to grounding in the 10% to 15% range, the majority put engineering costs at less than 5%. The understanding of transmission line grounding problems was considered to be satisfactory by 47% of the utilities surveyed and less than satisfactory by 10%. The remainder considered their understanding to be either good or very good. 11-3 11.5.2 MEASUREMENTS (SECTION II) The survey investigated field measurement practices of utilities in two areas; pre-design and post-construction. This section covered pre-design activities although some inconsistencies in the answers to this section suggest that this may not have been understood by most respondents. For example, only 36% answered "yes" to question 1, which asked whether or not field measurements were conducted prior to construction while the remaining questions in this section dealing with measurement techniques were answered by most respondents. Thus, these responses were interpreted in the context of all transmission line related measurements and not measurements related only to transmission line structure grounding design. Approximately one half of the responding utilities conduct soil resistivity measurements, primarily for the purpose of estimating the resistance of structure grounds or to categorize the soil structure in order to apply appropriate standard grounding arrangements. By far the most common technique used for resistivity measurement is the Wenner method using a Megger type instrument (null balance or direct reading). The problem of interpretation of earth resistivity measurements is handled using engineering judgment or design charts by all but two of the responding utilities who used more elaborate theoretical methods employing computers. Only 18% of the respondents used field data from geological surveys to assist in the interpretation of resistivity data. All but two of the utilities conducting earth resistivity measurements develop a uniform soil model from resistivity data. One utility uses a two-layer soil model and one uses an expotentially varying soil model. Eighty nine percent of the utilities responding to this section consider their field measurement procedures and interpretation techniques to be satisfactory or better, 11% consider them less than satisfactory. 17 utilities provided comments on the type of improvements required to make field measurements more useful and reliable. These comments are summarized in Section 11.6. 11.5.3 ENGINEERING (SECTION III) Seventy five percent of the responding utilities indicated that they carry out conceptual design calculations related to transmission line grounding, including lightning and 60 Hz fault conditions. Fourty eight percent of all utilities indicated that 60 Hz fault calculations were carried out. Computers or programmable calculators are used by 33% of those who carry out design calculation for transmission structure grounds. As suggested in the section dealing with measurements, most utilities use an equivalent uniform soil resistivity model in these calculations. Ground rods are the most common grounding electrode assumed in the calculations. The distribution of fault current between a transmission structure and skywires assumed by utilities in their 60 Hz fault calculations varied considerably. While 32% of those carrying out calculations assumed all 60 Hz current to flow in the structure, the remaining utilities assumed distributions as shown in Figure 11.2. Sixty four percent of all responding utilities indicated that design calculations were carried out related to lightning conditions on transmission lines. Most calculations include the soil resistivity and structure grounding leakage (dc) resistance in the calculations, while parameters such as soil permittivity, permeability and impulse breakdown gradient are seldom used. 11-4 5 5 4 (/l 3 hu. o cc LU CQ 2 2 1 . 1 X z 0 0 0 10 20 30 40 50 60 FAULT CURRENT IN TOWER (IN Figure 11.2 70 %. 80 90 100 ) Distribution of Fault Current Between Structure and skywires Keraunic levels are used by 95% of those utilities involved in lightning condition calculations, while the detailed electrical parameters of structures and structure footings are used by a relatively small number of utilities. As expected, all of the utilities who carry out engineering calculations for lightning conditions do so in order to estimate transmission line rates of failure. Estimating charts and empirical formulas are the principal tools used to carry out these calculations. The basic hypotheses used in rate of failure calculations for lightning conditions vary considerably except for soil structure, which is considered uniform by more than 90%. The structure impedance is ignored by 40% of these utilities; reflections and refractions at neighbouring structures are ignored by 47%; mutual coupling effects between phase and ground wires are ignored by 42%, while the increased mutual coupling effects of corona are ignored by 56% of the utilities in their lightning condition calculations. Most utilities who carry out 60 Hz and lightning calculations consider their methods to be satisfactory or better. Only two utilities considered these to be unsatisfactory. Fifteen utilities provided comments regarding improvements that could be made to make their calculation methods more accurate. These comments are summarized in Section 11.6. 11.5.4 SAFETY CONSIDERATIONS (SECTION IV) From the responses to this section, it is apparent that most utilities do not consider safety around transmission line structures to be a serious problem. About 75% of all utilities indicated that no investigations were conducted to determine public or personnel safety around transmission line structures for 60 Hz fault or lightning conditions. Hazard levels were calculated according to the IEEE Guide 80 standard in most cases. Sixteen percent of the respondents investigate the safety of structure grounding at all transmission structures, while another 13% conduct investigations in high exposure areas. Fourty six percent do not conduct any safety investigations related to structure grounding. In comparing the hazard levels within substations to those due to 60 Hz faults on transmission structures, out of the 68 utilities responding to this question, 73% believed the hazard level associated with transmission structure to be lower or significantly lower. Seven percent considered the hazard due to transmission structure faults to be greater, with the remainder considering the two hazard levels comparable. Seventy five percent of the responding utilities considered their approach to safety at transmission line structure to be acceptable and 15% considered it unacceptable. Ten percent did not reply to this question. 11-5 No utilities indicated that they were aware of any accident (fatal or otherwise) caused by 60 Hz faults on transmission structures. The consensus was that hazards associated with transmission structure grounding would remain constant or increase somewhat over the next ten years. The primary reasons given for increases in the hazard level were, higher voltage and fault capacity systems and increased exposure. Twenty utilities provided comments on corrections or improvements they believed necessary to increase safety around transmission structures. These comments are summarized in Section 11.6. 11.5.5 CONSTRUCTION / MEASUREMENTS / MAINTENANCE This section contained several questions relating to types and sizes of materials used in the construction of structure grounding electrodes. These results can be examined in the detailed response tabulation. Sixty four percent of the utilities indicated that field measurements are conducted at all structures after installation, to check the footing resistance. These measurements are, in almost all cases, used to initiate grounding system modifications if the resistance is inadequate. Thirty nine utilities provided estimates of the additional costs typically incurred by such modifications. These data are summarized in Figure 11.3. ADDITIONAL COST Figure 11.3 Additional Cost of Grounding One utility out of the 88 respondents indicated that staged fault tests were always conducted after line construction, while 85% indicated that staged fault tests are never conducted. Structure grounds are checked at some time by 75% of the utilities. Mechanical failures due to vandalism, accidental damage and corrosion were identified as the primary reasons for periodic or occasional grounding checks. Figure 11.4 illustrates the distribution of the estimated lifespan of ground rods and counterpoises provided by 53 utilities. The average of the estimate for each is around 30 years. 11-6 0 10 20 30 <t0 50 70 80 USEFUL LIFE (IN YEARS) Figure 11.4 Average Useful Life of Ground Conductors Most utilities keep records of the type and location of tripouts or protection system malfunctions. The principal cause of transmission line failure is identified as mechanical damage. As to the quality of their construction procedures, measurement techniques programs, most utilities considered these to be satisfactory or good. There noticeable shift downwards in perceived quality from construction to maintenance. From this and subsequent comments summarized in Section that maintenance is considered to be the most neglected area. and maintenance was, however, a measurement to 11.6, it is clear 11.5.6 TYPICAL DESIGN AND DATA (SECTION VI) This section surveyed the application of ground wires, grounding electrode configurations, desired ground resistances and failure rates. Ground wires, grounded at each structure and connected to the substation ground are employed by a large majority of utilities on all 100 kV class and higher systems. Two utilities, only partially install ground wires on 500 kV and 700 kV systems, and one utility makes no use of ground wires on 300 kV class systems. Twenty percent of the utilities partially install ground wires on 100 kV class systems. The results of the questions concerning structure grounding arrangements, desired ground resistance, and failure rates are summarized in the accompanying Figures 11.5, 11.6 and 11.7. The final question concerned the lightning failure rates of single versus double circuit lines and showed that a slight majority (58%) of those who answered this question did not consider double circuit line to be more failure prone. The utilities were grouped into those reporting maximum structure resistances of less than 30 ohms and greater than or equal to 30 ohms. The mean failure rates for each group in each voltage class were then determined with the results shown in Table 11.2. 11-7 VOLTAGE CLASS UTILITIES WITH MAX. TOWER RESISTANCE < 30 ^ No. of Failure Rate (Mean) Utilities Std. Dev. 100 kV 18 1.732 1.39 200 kV 17 1.37 300 kV 11 1.26 500 kV 6 Table 11.1 .404 UTILITIES WITH MAX. TOWER RESISTANCE > 30 £3 No. of Failure Rate (Mean) Utilities Std. Dev. 13 2.39 1.58 1.68 9 2.27 1.418 1.0 6 ■ 98 1.03 2 .27 .04 .38 Maximum Structure Ground Resistances The results of the comparisons for the 100 and 200 kV classes could be significant, suggesting that lower footing resistances will result in lower failure rates due to lightning. The values are close enough however, that other factors, not considered in this comparison, could distort the results. Keraunic levels, for example, have not been considered and a correlation between keraunic levels and average soil resistivities could affect the above results in either direction. In order to obtain additional information from the responses to the survey questionnaire, attempts were made to derive relationships between various transmission line grounding performance criteria and design criteria. Since no utilities were able to identify any safety problems related to transmission line grounding, only transmission line performance data were considered in the analysis. Initially, the utilities were grouped into high and low failure rate groups. A failure rate of one failure per 100 miles per year was used as the separation point. The responses to various engineering design related questions were then compared between the two failure rate groups but no meaningful relationships were observed. A detailed analysis was then undertaken of the relationship between structure resistances and lightning failure rates. This was investigated for each voltage class of transmission line. The mean values of maximum structure resistance for each class were compared between high and low failure rate companies. The results of this are shown in Table 11.2. VOLTAGE CLASS HIGH FAILURE RATE UTILITIES ( > FAILURE /100 MILES /YEAR) Max. Tower Footing No. of Res i stance £2 Utilities Mean Std. Dev. LOW FAILURE RATE UTILITIES ( < FAILURE /100 MILES /YEAR) No. of Utilities Max. Tower Footing Res i stance £2 Mean Std. Dev. 100 kV 6 107 194 8 26 31 200 kV 2 260 340 8 27 28 300 kV 6 23 24 3 15 10 500 kV 2 15 7 4 28 30 700 kV 1 10 0 1 25 0 Table 11.2 Transmission Line Lightning Failure Rates 11-8 Although 100 kV and 200 kV class mean values are significantly higher, the distribution is skewed considerably by one utility reporting a maximum value of 500 ohms. If these data are not considered, the values are very similar. However, it was difficult to draw any firm conclusions from this result. CONTINUOUS COUNTERPOISE RADIAL COUNTERPOISE nn n GROUND ROD ONLY TOWER FOUNDATION ONLY TT r~i-, a fl n- ■ n J□nD r—i czi J1 £=□D=l 300 kV ^ rTI 500 k\I 700 kV » 0 ^ ■____ ' 40 80 « 0 ■ 40 TYPE Figure 11.5 11' OF 80 -L- I « 0 GROUNDING t 40 I i 80 t ..g..,,! I 0 40 (IN % OF TOTAL) Grounding Arrangement 11-9 I i 80 DESIRED VALUE MINIMUM MEASURED MAX I MUM MEASURED i—i—1 0 1 100 kV o a LU m s: => ■z. 20 ’ —i 10- Tl n r-1 n . 200 kV UTILITIES » __i__i__i____i_ — o cc s s Z5 Z i “1 r-n— 20 * 10 ' 0 r jTh L ri - rffci 0 20 i+O 0 n& -T—T-T".. .<---- lAi M 4 ip RESISTANCE Figure 11.6 ——■cL-..-,------inLr* infa....... I 0 0 40 (IN OHMS) Transmission Structure Resistance AVERAGE MAXIMUM AVERAGE 100 kV NUMBER RATE MAX I MUM 200 kV 300 kV Figure 11.7 100 OF FAILURES Lightning Rate of Failures 11-10 11.6 SUMMARY OF COMMENTS RECEIVED NOTE : (n) gives the number of respondents who provided the comment. SECTION I - GENERAL 1-1 Which (7) (5) (4) (2) (2) Guide is Followed ? IEEE Guide No. 80 National Electric Safety Code Westinghouse Electrical T&D Reference Book EPRI Transmission Reference Book IEEE Practice for Grounding of Industry and Commercial Power Systems 1-2 Why Install Grounding Systems ? (5) To prevent lightning failures (4) To keep ground resistance below 10 ohms (4) To prevent structure damage (3) To minimize induced voltages on pipelines, telephone lines, etc. (2) To follow state regulation, overhead line general order 95 SECTION II - MEASUREMENTS II-4 Objectives of Taking Field Data ? (3) To help to decide the type & size of the grounding system (3) To verify the adequacy of the installed system (2) To assist in the design of the structure (1) To determine the necessity of putting in more grounding (1) To perform calculations to minimize the effect of transmission line on adjacent pipelines II-5 Methods Used to Measure Resistivity ? (3) Megger (3-point, 4-point) terminal test (2) Direct ohm-meter method of 2-terminal test II-6 Types of Instrument Used ? (3) Vibroground meter model 293 II-7 Interpretation of Measurements ? (2) To develop a two-layer model (1) Use "Tagg's" method to derive formula for ground resistivity determination 11-10 Improvements to Make Measurements & Interpretation More Reliable ? (4) A comprehensive set of standards on measurement techniques and interpretations (2) A better method to measure structure resistance (2) A better model for multi-layer nonuniform resistivities (1) Better equipment, more measurement (1) A formula of resistance as a function of seasonal soil moisture content 11-11 SECTION III - ENGINEERING III-l 60 Hz Fault Calculations Variables ? (2) Grounding conductor resistance HI-5 Distribution of Fault Current Between Tower Ground Electrode and Ground Wire ? (2) Percentage varies with line arrangements, structure spacing & ground conditions III-6 Lightning Calculation Variables ? (3) Shield angle (2) Air gap with steel (1) Flashover characteristics III-7 Methods of Determining Transmission Line Rate of Failure Due to Lightning ? (4) Methods referred to in EPRI Transmission Line Reference Book 345 kV & Above (3) Historic records (1) IEEE methods III-ll Improvements ? (4) Better guidelines (1) Better statistical data to indicate which design variables are more valuable than others (2) More related computer programs to be written SECTION IV - SAFETY CONSIDERATIONS IV-2 Source of Maximum Value of Tolerable Body Surge Current ? (3) IEEE Guide 80 (3) Dalziel's electrocution formula 1= 0.116/ VU (2) 5 mA, rule of thumb IV-6 Investigations of Safety of Structure Grounding ? (5) No investigations are conducted (3) Thorough & frequent checkings (2) Thorough checking plus special grounding in highly exposed areas (1) Only investigations where there are parallel pipeline exposures IV-8 Analyze Hazards During 60 Hz Fault or Not ? (2) Do not because it is believed that as long as ground resistance is held below 10 ohms, it is safe IV-9 Ever Analyze Special Structure Grounding Electrodes ? (3) Study variable counterpoise patterns (2) Study ring-shaped electrode IV-11 Improvements For Safety ? (2) Increase the engineering staff (2) Verification of formula & maximum current values that can be tolerated during faults (1) Some structure grounding should be designed specifically for public safety (1) Inform & educate the public about safety (1) Develop step & touch potential criteria 11-12 IV-13 Future Hazards Associated With Transmission Line Structure Grounding for Next Ten Years ? (21)Believe will go up because of: 1/ more line construction 2/ more intense land use which cause more exposures (3) Believe no real problem (1) Believe will go down due to better transmission grounding design and installations SECTION V - CONSTRUCTION / MEASUREMENT / MAINTENANCE V-2 Approximate Diameter of Conductor Used ? (2) AWG # 4 copper (1) Size depends on fault current V-4 Equipment Used to Install Structure Grounding ? (2) Rock drills & grout mixers (1) Compressor & air hammer V-6 Field Measurements Taken ? (4) Resistance of ground rod (1) Soil resistance V-9 Methods to Measure Structure Footing Leakage Resistance ? (3) 3-point method V-10 Equipment Used ? (1) BBC high frequency ground resistance tester (1) 3-probe null balance resistance testing instrument V-ll Purposes for Staged-Fault Tests ? (4) Relay coordination (1) Potential rise of gas pipeline under lines (1) Research purposes V-13 Reasons for Checking Transmission Structure Grounds ? (5) Better safety and line performance during lightning (4) To look for any damage caused by vandalism, agricultural activities, corrosion, etc. V-13 Methods Used for Checking Ground System ? (4) Visual inspections (2) Resistance measurements (1) Sample excavation V-17 Cause of Transmission Line Permanent Failure ? (5) Lightning storms (4) Trees (3) Vandalism (3) Accidents (2) Misoperations (2) Equipment failures V-19 Modification to Improve Grounding Performance ? (12) Better maintenance such as more inspection, better measurement techniques, better equipment, better trained .staff 11-13 SECTION VI - TYPICAL DESIGN AND DATA VI-1 Equipped With Overhead Ground Wire ? (1) Only apply to lines that perform poorly. 11.7 CONCLUSION The survey results overwhelmingly confirmed what was already suspected; North American power utilities do not follow any common, generally accepted transmission line grounding standards nor do they apply similar practices in measurement, design or construction procedures. The practices, needs and understanding of transmission line grounding by utilities varies widely. Conflicting concepts and opinions exist in almost every significant aspect of this important subject. It is hoped that this report will provide the framework for a coordinated and more unified approach to transmission line grounding problems. 11-14 APPENDIXES APPENDIX A COMPUTER PROGRAM LINPA FUNCTION LINPA computes the self and mutual impedances of phase and ground conductors of transmission lines. The program uses the equations developed by Carson, assuming a uniform soil structure. LINPA first calculates and prints the complete impedance matrix. Then, if requested, it performs bundle reductions and prints a reduced impedance matrix where the original group of conductors of the bundle are represented by a single equivalent conductor. REFERENCES Information on this program is given in Chapter 2. Carson's equations are described in references 1 and 3 of Chapter 2 and in several power engineering textbooks. A typical application of LINPA is described in Chapter 3. GENERAL LINPA was developed for use in an interactive mode. In this mode, the user is prompted to provide the required input data and answers in free-format. A typical interactive session is shown in Page A-5. The program can also be readily operated in batch mode. In this mode, the user must organize the data in a file prior to running the program. The batch mode is particularly advantageous when running several cases with small differences in the input data. The following guidelines and the interactive session shown in Page A-5 provide the necessary instructions to build an input data file for the program LINPA. GUIDELINES All values entered on a data line must be separated by a comma. Blanks are interpreted as a YES, zero or default answer, depending on the data expected by the program. LINPA accepts up to 8 different types of data lines. Occasionally, several data lines of the same type are accepted by the program. In this case, the user must enter a line consisting of the world END as the last entry to instruct the program that the series of data lines of the current type has ended. A typical input data file is shown in Page A-4. This file can be prepared based on the following guidelines: A-1 Data Line 1 : Start This line must necessarily be a YES to initiate execution of the program. A NO will stop the program. Data Line 2 : Comment Enter any alphanumeric character to provide comments and general references. These lines are not processed by the program but simply reprinted at the top of the output printout. Up to four lines of comments may be entered. If less than four comment lines are entered, the comments must be terminated with a line consisting of the word END. If the character is among the characters entered in a comment line, all the information to the right of this character is ignored. This is a useful feature (for both interactive and batch modes) which allows the user to include additional comments in the input data which will not appear in the printout. Data Line 3 : Run Identification Enter a four digit identification code which will be printed at the top right-hand side of each page of printout. Data Line 4 : System of Units If the metric system is to be used, provide a data line beginning with a YES. If the British system of units is requested, two data lines are needed; line 1 is a NO and line 2 is a YES. Data Line 5 : Bundle Reduction Enter YES if the program is to reduce all the conductors of a bundle into an equivalent single conductor. Enter NO if bundle reduction is not required. The conductors in a bundle have the same phase number. Data Line 6 : System Frequency Enter the frequency (Hz) at which the impedance matrix is to be calculated. A zero frequency is interpreted by the program as direct current. Data Line 7 : Earth Resistivity This is the average resistivity value of the soil along the length of the transmission line consi dered. This value should always be expressed in ohm-meters regardless of the system of units selected. A-2 Data Line 8 : Conductor Data Provide, in one line, the characteristic values of each conductor. The characteristic values, separated by commas, are illustrated in the following line: PHA, CODE, X, Y, RADIUS, XXX, RAC, NO, ORAD, RCOR where PHA is the phase number of the conductor. Note that this phase number does not play any role other than grouping the conductors into a number of bundles to be replaced by eguivalent single conductors. For example, the two ground wires of a transmission line can be assigned the phase number 4 (or 7 for a double circuit line) in order to reduce them to the eguivalent ground wire needed for the program PATHS. CODE represent, the digits 0 and 1. If 0 is selected, LINPA interprets the following XXX entry as being the GMR of the conductor. If 1 is used, XXX is interpreted as being the conductor inductive reactance at one foot spacing. X is the horizontal distance of the conductor as measured from an arbitrary origin on the surface of earth. This distance is in m or ft depending on the units selected. Y is the average height above ground of the conductor expressed in m or ft. RADIUS is the conductor radius in m or ft. XXX is the geometric mean radius (GMR) of the conductor (in m or ft) or its inductive reactance at one foot spacing (in ohms/km or ohms/mi), depending on the value of code. RAC is the 60 Hz ac internal resistance of the conductor in ohms/km or ohms/mi. NO is the number of outer strands of the conductor. This value system frequency (data line 6) is greater or equal to 2,000 Hz. is needed only if the ORAD is the radius of the outer strands (m or ft). This value frequency is greater than or equal to 2,000 Hz. is needed only if the RCORE is the overall radius of the conductor core (m or ft). If all the strands are of the same material, then RCORE = ORAD. If the conductor is a solid conductor then RCORE = RADIUS. This value is needed only if the frequency is greater or equal to 2,000 Hz. The user should provide as many data lines of type 8 as there are transmission line conductors, up to a maximum of 25 which is the limit of the released version of LINPA. A data line with END in the first columns must be the last line entered. A-3 INPUT DATA FILE FOR BATCH MODE PROCESSING BABE TEST FILE (APPENDIX A) EXAMPLE FOR CHAPTER 3 TYPICAL LINE 345 KV DETROIT EDISON CACSR 54/7 #954 EX1 N Y Y 60 150 l,Or-19.33333,48.*.04983,.0404*.0998 1*0*-20.83333*48.*.04983 *.0404 *.0998 2*0,-26.83333*69.5*.04983*.0404*.0998 2*0*-28 * 33333 *69.5*.04983 *.0404 *.0998 3,0*-20.33333,94.*.04983*.0404*.0998 3.0, -21.83333,94.,.04983*.0404,.0998 4.0, -12*0,109.,.02121,0.00276,1.437 4.0. 12.0.109...02121.0.00276.1.437 5.0. 19.33333.48...04983..0404..0998 5.0. 20.83333.48...04983..0404..0998 6.0. 26.83333.69.5..04983..0404..0998 6.0. 28.33333.69.5*.04983,.0404,.0993 7.0. 20.33333.94...04983..0404..0998 7.0. 21.83333.94...04983..0404..0998 END A-4 *2 19*10) TYPICAL INTERACTIVE DIALOGUE AND COMPUTER PROGRAM PRINTOUT PR06RM LINPA THIS PROSRAH CALCULATES THE IHPEBAHCES OF TRANSMISSION LM CONBUCTORS (PHASE LINE PARAMETERS) NOTES 1- RESPONO TO OJESTIONS BY YES OR NO OR* INCH APPRIMIATEt BY TYPING THE REQUIRED DATA. 2- Y* YE* YES* BLANK CHARACTERS AND RETURN CARRIAGE ARE INTERPRETED AS YES, OTHER CHARACTERS ARE INTERPRETED AS NO, 3- 0* 0,* BLANK CHARACTERS OR A RETURN CARRIAGE ARE INTERPRETED AS A ZERO VALUE OR* HHEN APPLICABLE* ARE REPLACED BY DEFAULT VALUES, START ? yes ALL CONDUCTORS BELONGING TO THE SAME BUNDLE ARE* BY CONVENTION, CONSIDERED AS FORMING ONE IPHASE*. THEREFORE* IF THE BUNDLE REDUCTION OPTION IS ENABLED, THE PROGRAM HILL REDUCE THE BUNDLE INTO ONE EQUIVALENT SINGLE CONDUCTOR. FOR EXAMPLE ,* - IF SEVERAL OVERHEAD GROUND-HIRES ARE TO BE REDUCED INTO ONE EQUIVALENT CONDUCTOR DECLARE EACH GROUND-HIRE AS PART OF THE SAME PHASE* I.E,* PHASE #4 FOR A SINGLE-CIRCUIT 3-PHASE LINE OR PHASE * 7 FOR A DOUBLE CIRCUIT 3-PHASE LINE - IF YOU DISH TO COMBINE PHASE A1 AND PHASE A2 OF A DOUBLE CIRCUIT LINE DECLARE THE CONDUCTORS OF BOTH BUNDLES AS PART OF THE SAME FICTITIOUS PHASE * N ENTER COMMENT LINES. (MAXIMUM OF 4) TERMINATE HITH END AT THE BEGINNING OF A NEU LINE, file (appendix A) for chapter 3 typical line 345 KV Detroit Edison (ACSR 54/7 #954,2 19*10) ?? ? bsse exaepletest ? ENTER RUN IDENTIFICATION (4 DIGITS), ? exl SYSTEM OF UNITS - METRIC ? ? no - BRITISH ? ? yes DO YOU HANT TO REDUCE THE CONDUCTOR BUNDLES INTO EQUIVALENT SINGLE CONDUCTORS ? ? yes ENTER THE POMER SYSTEM FREQUENCY (HERTZ) ? m ENTER EARTH RESISTIVITY (OHMS-HETER) ? 150 A-5 CONDUCTOR'S DATA (MAX. OF 25 CONDUCTORS) YOU CAN ENTER THE GHR OR THE INDUCTIVE REACTANCE AT ONE FOOT SPACING (60 HERTZ), GMR IS USED WHEN THE CODE (SEE BELOW) IS 0 OR BLANK* OTHERWISE THE REACTANCE XI AT ONE FOOT SPACING (60 HERTZ) IS USED, ENTER THE PHASE NUMBER* THE CODE* THE X AND T COORDINATES (FT) THE RADIUS (FT)* THE GMR (FT) (OR XI(OHMS/MILE)) AND THE AC RESISTANCE (OHHS/HILE) OF THE CONDUCTORS, TERMINATE WITH END AT THE BEGINNING OF A NEW LINE, EXAMPLE : PHA*CGDE*X*Y*RADIUS*GMR*RAC ? 1*0*-19.33333*48.*,04983*,0404*.0993 ? l»6*-20,83333*48,*,04983*.0404*,0998 ? 2*0*-26,83333 *69♦5 * *04983 *,0404*,0998 ? 2*0*-28,33333*69,5.,04983.,0404.,0998 ? 3.0*-20,33333*94,*.04983*,0404*.0998 ? 3*0*-21.83333.94,.,04983*.0404.,0998 ? 4*0*-12.0*109,.,02121.0,00276.1.43? ? 4.0.12,0.109,.,02121.0,00276.1,437 ? 5.0*19,33333*48,*.04983?,0404.,0998 5.0.20,83333.48,.,04983.,0404.,0998 6*0.26,83333*69,5.,04983.,0404*,0993 6*0*28,33333.69,5.,04983.,0404.,0998 7.0.20,33333.94,.,04983.,0404.,0998 7.0.21,83333.94..,04983.,0404..0998 ? END ? ? ? ? ? A-6 LINPA PAGE RUN J EX1 comments XBASE TEST FILE (APPENDIX A) mt &wb 3 IDETSOIT EDISON (ACSR 54/7 #954 RUN IDENTIFICATION UNITS SYSTEM 1 mmmmmttnmmmtmm t i t 19110) COMMENTS * 5 EX1 1 BRITISH BUNDLE REDUCTION : YES CALCULATION METHOD : CARSON SYSTEM FREQUENCY : 60,00 HERTZ SOIL RESISTIVITY l 150.00 OHMS-METER NUMBER OF CONDUCTORS S 14 LINPA PAGE 1 2 RUN : EX1 CONDUCTORS DATA AS ENTERED BY USER t-------------------------------------- 1 CONDUCTOR PHASE NUMBER NUMBER 1 2 3 4 5 h 1 8 9 10 11 12 13 14 1 1 2 2 3 3 4 4 5 5 6 6 7 7 X FEET Y FEET -19.33 -20.83 -26,83 -28.33 -20.33 -21.83 -12,00 12,00 19,33 48,00 48,00 69,50 69,50 94.00 94,00 109,00 109,00 48,00 48,00 69,50 69,50 94,00 94,00 3A 28.33 20,33 21.83 RADIUS FEET GMR FEET .4983E-01 ,4983E-01 .4983E-01 ,4983E-0i .4983E-01 .4983E-01 .2121E-01 .2121E-01 .4983E-01 .4983E-01 .4983E-01 .4983E-01 .4983E-01 ,4983E-01 .4040E-01 .4040E-01 .4040E-01 .4040E-01 .4040E-01 .4040E-01 .2760E-02 .2760E-02 .4040E-01 .4040E-01 ,4040E-01 .4040E-01 .4040E-01 .4040E-01 END OF INPUT DATA A-7 RAC OHHS/HILE .9980E-01 •9980E-01 .9980E-01 ,9980E-0i .9980E-01 .9980E-01 .1437EI01 .1437E+01 ,9980E-01 .9980E-01 .9980E-01 .9980E-01 .9980E-01 .9980E-01 LINPA PAGE I 3 RUN S EX1 CONDUCTORS DATA AS STORED BY PROGRAM --- ¥ <—-— CONDUCTOR PHASE NUMBER NUMBER 1 2 3 4 5 6 7 1 1 2 2 3 3 8 4 4 9 5 10 5 11 12 13 14 6 & 7 7 X METERS -5,89 -6,35 -8,18 -8,64 -6,20 -6.65 -3,66 3.66 5,89 6.35 8.18 8.64 6.20 6.65 Y METERS 14.63 14.63 21,18 21,18 28,65 28,65 33.22 33.22 14,63 14.63 21.18 21.18 28.65 28,65 RADIUS METERS PERMEABILITY RELATIVE RAC OHMS/KM ,1519£-01 .1519E-01 .1519E-01 .1519E-01 .1519E-01 .1519E-01 .6465E-02 .6465E-02 .839IETOO .6203E-01 .6203E-01 .6203E-01 .6203E-01 .6203E-01 •6203E-01 ,8931ET00 .8931EW0 .6203E-01 .6203E-01 .6203E-01 .6203E-01 .6203E-01 .6203E-01 ,15191-01 ,1519E-01 .1519E-01 ,1519E-01 .1519E-01 .1519E-01 .8391E+00 .8391E+00 .8391ET00 .8391E+00 •8391E+00 .8157E+01 .8157E+01 ,8391EW .8391E+00 .8391E+00 .8391E+00 .8391ET00 .8391E+00 END OF INPUT DATA A-8 LINPA PAGE 5 RUN I EX1 DETAILED LINE PARAHETERS AT IHPEDANCES FIRST 4 60,00 HZ - OHHS/HILE VALUE IS RESISTANCE. SECOND VALUE IS REACTANCE COND, NUHB, 4 COND. NUHB, i .192 1.395 2 .092 ,941 .192 1.395 3 ,092 .612 ,092 ,614 .191 1.396 4 ,092 .609 .092 .612 .091 .942 .191 1,396 5 .091 .527 .091 .527 ,091 ,600 .091 ,598 6 .091 ,527 ,091 ,527 ,091 ,602 ,091 ,600 7 .091 ,492 .091 .492 .090 ,539 ,090 ,537 8 ,091 ,479 ,091 ,478 .090 .506 .090 ,503 9 ,092 .547 .092 ,542 ,092 ,514 ,092 ,511 10 .092 .542 .092 ,538 .092 ,511 .092 .508 11 .092 ,514 ,092 .511 .091 .508 ,091 ,505 12 .092 .511 .092 .508 ,091 .505 ,091 .502 13 .091 .493 .091 .491 ,091 .510 ,091 .507 14 ,091 ,491 .091 ,489 .091 .507 ,091 ,504 COND. NUHB. 6 5 8 7 MB. MB. 5 ,190 1.398 6 ,090 .944 .190 1,398 7 ,090 .649 .090 ,643 1.526 1,859 8 .090 .560 .090 ,555 ,089 .608 1.526 1.859 9 .091 .493 .091 ,491 ,091 .479 .091 ,492 10 .091 .491 .091 ,489 .091 ,478 ,091 .492 11 ,091 ,510 .091 .507 .090 ,506 .090 .539 12 ,091 ,507 .091 ,504 ,090 ,503 ,090 ,537 13 .090 ,543 .090 ,539 .090 ,560 ,090 ,649 14 .090 ,539 .090 .535 .090 ,555 .090 ,643 A-9 CON!), COND, NUHB NUHB, 11 10 9 12 COND. NUHB. 9 .192 1.395 10 .092 .941 .192 1.395 11 .092 .012 .092 .614 ,191 1.396 12 .092 ,609 .092 .612 ,091 ,942 ,191 1,396 13 .091 .527 ,091 .527 ,091 .600 .091 .598 14 .091 .527 .091 .527 .091 ,602 ,091 ,600 COND. 14 13 NUHB, 1§ggg 1 13 ,190 1.398 14 .090 ,944 .190 1,398 LINPA RUN PHASE LINE PAGE ! 5 l EX1 PARAHETERS IHPEDANCES - OHHS/HILE FIRST VALUE IS RESISTANCE) SECOND VALUE COND. NUHB. IS REACTANCE 4 3 1 COND. NUHB. 1 ,142 1,168 2 .092 .611 ,141 1.169 3 ,091 ,527 .091 ,599 .141 1.169 4 .091 ,485 ,090 .521 ,090 .602 ,808 1.234 5 ,092 ,542 ,092 ,511 ,091 .492 .091 .485 6 .092 .511 ,091 ,505 ,090 ,508 ,090 .521 7 ,091 ,492 .090 ,508 .088 ,540 .090 ,602 COND. NUHB. 6 5 7 COND. NUHB. 5 ,142 1.168 6 ,092 .611 ,141 1,169 7 .091 ,527 .091 ,599 END OF PRGGRAH LINPA tmtunnnntnn A-l o .141 1.169 APPENDIX B COMPUTER PROGRAM RESIST FUNCTION This program determines an equivalent two-layer earth model from the measured apparent resistivity data. The resistivity values must have been measured using the equally-spaced four probe or Wenner method. The equivalent earth model is characterized by the thickness of the first layer and by the resistivity values of the upper and lower layers of soil. REFERENCES Information on this program is given in Chapter 2. The analytical theories on which RESIST is based are described in Chapter 4. A typical application of this program is given in Chapter 3. Additional examples are included in Chapter 10. GENERAL RESIST was developed for use in an interactive mode. In this mode, the user is prompted to provide the required input data and answers in free-format. A typical interactive session is shown in Page B-5. The program can also be readily operated in batch mode. In this mode, the user must organize the data in a file prior to running the program. The batch mode is particularly advantageous when running several cases with small differences in the input data. The following guidelines and the interactive session shown in Page B-5 provide the necessary instructions to build an input data file for the program RESIST. GUIDELINES All values entered on a data line must be separated by a comma. Blanks are interpreted as a YES, zero or default answer, depending on the data expected by the program. RESIST accepts up to 9 different types of data lines. Occasionally, several data lines of the same type are accepted by the program. In this case, the user must enter a line consisting of the world END as the last entry to instruct the program that the series of data lines of the current type has ended. A typical input data file is shown in Page B-4. This file can be prepared based on the following guidelines: B-l Data Line 1 : Start This line must necessarily be a YES to initiate execution of the program. A NO will stop the program. Data Line 2 : Comment Enter any alphanumeric character to provide comments and general references. These lines are not processed by the program but simply reprinted at the top of the output printout. Up to four lines of comments may be entered. If less than four comment lines are entered, the comments must be terminated with a line consisting of the word END. If the character is among the characters entered in a comment line, all the information to the right of this character is ignored. This is a useful feature (for both interactive and batch modes) which allows the user to include additional comments in the input data which will not appear in the printout. Data Line 3 : Run Identification Enter a four digit identification code which will be printed at the top right-hand side of each page of printout. Data Line 4 : System of Units If the metric system is to be used, provide a data line beginning with a YES. If the British system of units is requested, two data lines are needed; line I is a NO and line 2 is a YES. Data Line 5 : Type of Terminal Used Plots of measured and computed earth resistivities are automatically produced by the program for display on a terminal. If the terminal is a video screen, only one character at a time can be displayed in a specific location. This is not the case of printers and hard-copy terminals which allow overprinting. The program uses two different routines to produce plots, depending on the type of terminal. The hard-copy terminal routine produce the most accurate plots. However, the other routine can be used on a video or hard-copy terminal. This line should be a YES for a video terminal and a NO for a hard-copy terminal. Data line 6 : Maximum Probe Spacing Once the equivalent two-layer earth model has been determined, RESIST computes the apparent resistivity values which would have been obtained along the resistivity traverse if the earth structure was identical to the computed model. A total of 50 points are computed along the traverse. The maximum probe spacing of the array (in m or ft) must be defined by the user in this data line. B-2 If zeroes or blanks are entered in this line, RESIST assumes that the maximum probe spacing is 1.5 times the largest probe spacing used in the field measurements. This choice leads to a balanced appearance of the measured and computed curves of the plot produced by RESIST. Data Line 7 : Measured Apparent Values The measured results may be entered as apparent resistances or as apparent resistivities. A NO in this data line means apparent resistivities have been entered. A YES instructs the program that apparent resistances have been entered. RESIST will then calculate apparent resistivity values according to the equation: p = 27TaR where P is the resistivity, R the resistance and a the spacing. The above equation is only valid when the probe depth is approximately 1/10 or less of the probe spacing. In cases where this is not so or other influencing factors are known to be present, the user should determine apparent resistivity values using a more appropriate equation. Data Line 8 : Measured Values In this line, the user enters the probe spacing (in m or ft) and measured apparent resistance (in ohms) or apparent resistivity (in ohm-meters). The two values should be separated by a comma. The user must enter one data line for each pair of measured values. As explained, a line beginning with END instructs the program that all measured points have been entered. Date Line 9 : Accuracy Values Two accuracy values (in p.u.) are expected by the program in this data line. If the user does not specify, RESIST selects two default values. The first value represents the optimization accuracy of the earth model search algorithm. A value of 0.001 is often a very good choice. However, two runs may sometimes be necessary to achieve an optimum fit. In the first run, a large value such as 0.01 or the default value should be selected to speed-up the computation. The computed and measured curves should then be examined. If the fit is reasonable or if a better match is doubtful because the real soil is a very distinct three or four layer model, then a second run should not be made. It is recommended however, that the user try different accuracy values to obtain a good feel for the sensitivity of the program to changes in specified optimization accuracy. The second accuracy value is used to determine the end of the computation of the series terms in the development of two-layer models. The default value of 0.01 (1%) is an appropriate choice. B-3 INPUT DATA FILE FOR BATCH MODE PROCESSING YES BASE TEST FILE <APPENDIX B> EXAMPLE FOR CHAPTER 3 END EX1 YES YES O. NO 2.5*320 5.0*245. 7.5*182. lO.*162 12.5*168. 15.* 152. END 0.00001*0.001 TYPICAL INTERACTIVE DIALOGUE AND COMPUTER PROGRAM PRINTOUT PMHUft RESIST ttttutttutt AHALYSIS ASD INTERPRETATION OF APPARENT RESISTIVITY HEASURENENT AROUND TRANSMISSION LINE STRUCTURES NOTES I 1-RESP0ND TO SUESTIONS BY YES m NO OR? WHEN APPROPRIATE! BY TYPING THE REQUIRED DATA, 2-Y» YE? YES? BLANK CHARACTERS AND RETURN CARRIAGE ARE INTERPRETED AS YES, OTHER CHARACTERS ARE INTERPRETED AS NO, 3-0? 0,? BLANK CHARACTERS OR A RETURN CARRIAGE ARE INTERPRETED AS A ZERO VALUE OR? HHEN APPLICABLE? ARE REPLACED BY DEFAULT VALUES, START ? ? aes -“3 ENTER COMMENT LINES (MAX, OF 4) TERMINATE HITH END AT THE BEGINNING OF A NEW LINE, * — base test file (appendix B) exatple for chapter 3 ENTER RUN IDENTIFICATION (FOUR DIGITS) ? exl UNITS USED -METRIC SYSTEM ? ? yes VIDEO TERMINAL OR HARD-COPY TERMINAL (PRINTER) FOR OUTPUT ? VIDEO ? t yes ENTER MAXIMUM PROBE SPACING (METERS OR FEET) FOR THE CALCULATED APPARENT RESISTIVITY PROFILE* DEFAULTS 1,5 TIMES MAXIMUM SPACING OF THE MEASURED PROFILE, YOU MAY ENTER 1- THE APPARENT MEASURED RESISTANCE (V/D? 2- THE APPARENT RESISTIVITY VALUES OR APPARENT RESISTANCE T ? no ENTER YOUR FIELD DATA RESULTS IN SEQUENCE, ONE LINE FOR EACH MEASUREMENT AT A GIVEN SPACING, TERMINATE HITH END AT THE BEGINNING OF A NEW LINE, EXAMPLE S SPACING (M OR FT)? MEASURED VALUE (OHMS OR OHMS-METER) ? 2,5?320 ? ? ? ? ? 5,0?433,0 5*0?245. 7,5?182, 10.?162 12.5?168. 15*?152* ? END ENTER * OPTIMIZATION ACCURACY? AND SUM OF SERIES TERMS ACCURACY. (2 LAYER SOIL) (PER UNIT VALUES, DEFAULT VALUES ARE 0,01 AND 0.01) EXAMPLE l 0,005?0,001 0,00001?0,001 B-5 RESIST PAGE t 1 RUN : EX1 mmmmmmmmn comhents nmnmmmmtmn BASE TEST FILE (APPENDIX B) EXAMPLE FOR CHAPTER 3 ttWtttttttttttttttMtttt COMMENTS tttttttMtttttttttttttttt RUN IDENTIFICATION t EX1 SPACING IN METERS OPTIMIZATION ACCURACY ! .000010 (P.U.) SUM OF SERIES TERMS ACCURACY 5 .001000 (P.U.) BEST POSSIBLE FIT OCCUREO BEFORE REQUIRED ACCURACY HAS OBTAINED TRYING TO CONTINUE FIT PROCESS MAY LEAD TO DIVERGENCE ACTUAL ACCURACY = .009148 BEST ACCURACY = .001140 REACHED ON ITERATION 188 RESIST PAGE I 2 RUN 1 EX1 MEASURED APPARENT VALUES SPACING (M OR FT) 2.5000 5.0000 7.5000 10.0000 12,5000 15,0000 RESISTANCE RESISTIVITY OHMS-METER OHMS 320,0000 20,3718 245.0000 7,7986 182.0000 3,8622 162.0000 2,5783 168.0000 2.1390 152,0000 1,6128 RESIST PAGE ! 3 RUN ! EX1 OPTIMIZATION STATUS NUMBER OF ITERATIONS ? AVERAGE UMBER OF SERIES TERMS USED FOR TWO LAYER SOIL i OPTIMIZATION ACCURACY 1 319 45 ,1140 % B-6 RESIST PAGE J 4 RW I EX1 RESULTS COHPUTATIOH TOP LAYER RESISTIVITY 383,4982 OHMS-METER BOTTOM LAYER RESISTIVITY 147.6571 OHMS-METER REFLECTION FACTOR -.4440 (P.U.) TOP LAYER THICKNESS 2.5626 METERS RESIST PAGE I 5 RUN ! EX1 COMPUTED APPARENT RESISTIVITY PROBE SPACING (METERS) ,0100 ,9000 1.8000 2,7000 3,6000 4.5000 5,4000 fcSR 8,1000 9,0000 9.9000 10,8000 11,7000 12,6000 13,5000 14,4000 15,3000 16,2000 17,1000 18.0000 18,9000 19,8000 20.7000 21,6000 APPARENT RESISTIVITY (OHMS-METER) 383,4982 378,8418 355,4416 318,8212 280.9366 248,6495 223,5756 204,8539 191.1726 181,1541 173,8834 168,5325 164,5478 161.6009 159.3079 157.5274 156,1245 155.0033 154,0953 153,3506 152,7327 152,2147 151,7761 151.4015 151,0791 PROBE SPACING (METERS) ,4500 1,3500 2,2500 3.1500 4.0500 4,9500 5,3500 8,5500 9,4500 10.3500 11,2500 12,1500 13.0500 13.9500 14,8500 15,7500 16,6500 17,5500 18,4500 19.3500 20,2500 21,1500 22,0500 APPARENT RESISTIVITY (OHMS-METER) 382.8643 369,5771 337,9856 299.4948 263,9175 235,2636 213,4963 197,4698 185.7408 177,2362 171,0070 166,3970 162,9967 160.3800 158,3631 156 97855 155*5335 154,5262 153.7051 153,0277 152,4626 151,9865 151.5816 151,2344 150,9345 RESIST PAGE 1 6 RUN : EX1 COMPARISONS SPACING CALCULATED APPARENT MEASURED APPARENT DISCREPANCY (METERS) RESISTIVITY(OHMS-M) RESISTIVITY(OHMS-M) (PERCENT) 2.500 5,000 7.500 10,000 12,500 15,000 327,4372 233.8807 187,4388 168,0307 159,5344 155,3505 320,0000 245,0000 182,0000 162,0000 168.0000 152,0000 2,32 -4,54 2,99 3.72 -5,04 2.20 AVERAGE DISCREPANCY BETWEEN MEASURED AND CALCULATED RESISTIVITIES .28 PERCENT B-7 RESIST PAGE I 8 RUN 5 EX1 APPARENT RESISTIVITY OHHS-HETER 0< 39* 78* 0*0000-1................................. . I I I 117, 156. 195, .1______I______ I— 4,6000-1 I I I I I I I I I 9,2000-1 I I I I I I I I I 13.8000-1 I I I I I I I I I 18,4000-1 I I I I I I I I I , KEASURED RESISTIVITY f CONFUTED RESISTIVITY APPARENT RESISTIVITY END OF RESIST 234, 273. 312, 351, ...I______I______ I______ I— APPENDIX C COMPUTER PROGRAM GTOWER FUNCTION GTOWER computes the ground resistance of a transmission line structure and the earth surface potential profiles at any point around the base of the tower. The program can be used to analyze any structure grounding system made of cylindrical horizontal and vertical conductors buried in a uniform or two-layer earth. REFERENCES Information on this program is given in Chapter 2. The theoretical basis of GTOWER is described in Chapter 5. A typical application of GTOWER is shown in Chapter 3. Other examples are examined in Chapter 10. GENERAL GTOWER was developed for use in an interactive mode. In this mode, the user is prompted to provide the required input data and answers in free-format. A typical interactive session is shown in Page C-6. The program can also be readily operated in batch mode. In this mode, the user must organize the data in a file prior to running the program. The batch mode is particularly advantageous when running several cases with small differences in the input data. The following guidelines and the interactive session shown in Page C-6 provide the necessary instructions to build an input data file for the program GTOWER. GUIDELINES All values entered on a data line must be separated by a comma. Blanks are interpreted as a YES, zero or default answer, depending on the data expected by the program. GTOWER accepts up to 10 different types of data lines. Occasionally, several data lines of the same type are accepted by the program. In this case, the user must enter a line consisting of the world END as the last entry to instruct the program that the series of data lines of the current type has ended. A typical input data file is shown in Page C-5. This file can be prepared based on the following guidelines: C-1 Data Line 1 : Start This line must necessarily be a YES to initiate execution of the program. A NO will stop the program. Data Line 2 : Comment Enter any alphanumeric character to provide comments and general references. These lines are not processed by the program but simply reprinted at the top of the output printout. Up to four lines of comments may be entered. If less than four comment lines are entered, the comments must be terminated with a line consisting of the word END. If the character is among the characters entered in a comment line, all the information to the right of this character is ignored. This is a useful feature (for both interactive and batch modes) which allows the user to include additional comments in the input data which will not appear in the printout. Data Line 3 : Run Identification Enter a four digit identification code which will be printed at the top right-hand side of each page of printout. Data Line 4 : System of Units If the metric system is to be used, provide a data line beginning with a YES. If the British system of units is requested, two data lines are needed; line 1 is a NO and line 2 is a YES. Data Line 5 : Symmetry Generally, a transmission line structure is symmetrical. If so, only ground conductors in the positive quadrant need be inputted if the coordinate system is centered at the base of the structure. Program GTOWER does not accept as input, ground conductors located outside of the positive quadrant of the coordinate system which has its xoy plane on the surface with the oz axis directed toward the center of earth. This is in no way a limitation of the program, since the user can always locate the coordinate system far enough from the grounding system so as to include the entire grounding system in the positive quadrant. A line beginning with YES instructs the program that the grounding system is symmetrical and that the ox and oy axes are the symmetrical lines. The origin of the coordinate system is also a point of symmetry. A NO entry simply means that the grounding system is in the positive quadrant of the coordinate system. C-2 Data Line 6 : Type of Terminal Used Plots of the earth surface potential profiles are automatically produced by the program on a terminal. If the terminal is a video screen, only one character at a time can be displayed in a specific location. This is not the case of printers and hard-copy terminals which allow overprinting. The program uses two different routines to produce plots, depending on the type of terminal. The hard-copy terminal routine produce the most accurate plots. However, the other routine can be used on a video or a hard-copy terminal. Enter a YES to select video terminal and a NO to select the hard-copy terminal. Data Line 7 : Soil Data This line may contain up to three values, separated by commas. The first two values are the upper and lower layer earth resistivities of the earth respectively (in ohm-meters). The third value is the upper layer height (in m or ft). If the earth is uniform, only the first resistivity value need be entered. If either the lower layer resistivity or upper layer height is zero, then uniform earth is assumed. Data Line 8 : Fault Current This line provides the fault current flowing to earth through the grounding system (in amperes). Data Line 9 : Ground Conductors The user must enter one line per grounding system conductor or one line per positive quadrant conductor if the system is symmetrical. If a ground conductor is curvilinear, it must first be broken down into appropriate linear segments. The ground conductors are entered in two steps. In the first step, all horizontal conductors are entered. The last entry must be followed by an END line. If no horizontal conductors exist, the END line must still be entered. In the next step, the vertical ground elements are entered. An END line must be the last entry whether or not there are vertical elements. At least one, and no more than 67 conductors can be specified. The radius of a vertical conductor can not exceed 2 m (6.5 ft) and that of a horizontal conductor must be equal or less to 0.5 m ( 1.6 ft). Horizontal conductors must be completely buried in earth. Horizontal Conductors The values in the following line are expected by GTOWER, separated by commas: XO, YO, XE, YE, RAD, DEP C-3 where XO and YO are the X and Y coordinates of one end (origin) of the conductor (m or ft). XE and YE are the X and Y coordinates of the other end (extremity) of the conductor (m or ft). RAD is the radius of the conductor (m or ft). DEP is the depth of the conductor (m or ft). Vertical Conductors The following data line must be entered to identify vertical conductors: X, Y, RAD, LEN, TED where Xand Y are the coordinates of the vertical conductor (m or ft). RAD is the radius of the conductor (m or ft) LEN is the length of the conductor (m or ft) TED is the depth of the upper end of the conductor (m or ft). Data Line 10 : Potential Profiles The user may select up to ten profiles along which earth surface potentials will be computed at specific intervals. Each profile is defined as follows: XO, YO, XE, YE, SPA where XO and YO are the X and Y coordinates of the origin of the profile (m or XE and YE are the X and Y coordinates of the extremity of theprofile ft). (m or ft). SPA is the distance between two consecutive points of the profile (m or ft). A line beginning with END instructs GTOWER that no more profiles are to be calculated. Note that the profile points may be in any quadrant of the coordinate system. C-k INPUT DATA FILE FOR BATCH MODE PROCESSING Y BASE TEST FILE < APPENDIX C) EXAMPLE FOR CHAPTER 3 TOWER GROUNDING PLUS ONE RING END EX2 N Y Y Y 383.5*147.&6»8*405 lOOO 13.5*12.3*12.3*13.5*0.04 » 5 * 4 13.7*12.5*12.5*13.7*0.04*5.4 13.9*12.7*12.7*13.9*0.04*5.4 14.1*12.9*12.9*14.1*0.04*5.4 14.3*13.1*13.1*14.3*0.04*5.4 14.5*13.3*13.3*14.5*0.04*5.4 0.0*18.5*18.5*18.5*0.04*0.75 18.5*0.0*18.5*18.5*0.04*0.75 END 14.* 14.*0.04*5.36*0.04 END 14.0*0.0*14.0*30.0*0.5 END C-5 TYPICAL INTERACTIVE DIALOGUE AND COMPUTER PROGRAM PRINTOUT PROGRAM GTOWER tttttttttttttt THIS PROGRAM CALCULATES THE PERFORMANCE OF TRANSMISSION LINE GROUNDING SYSTEMS DURING POWER FREQUENCY FAULTS! - RESISTANCE OF GROWING SYSTEM - EARTH SURFACE POTENTIALS (STEP AND TOUCH VOLTAGES) NOTES ! 1-RESPOND TO QUESTIONS BY YES OR NO ORf WHEN APPROPRIATEf BY TYPING THE REQUIRED DATA, 2-Yf YEf YESf BLANK CHARACTERS AND RETURN CARRIAGE ARE INTERPRETED AS YES. OTHER CHARACTERS ARE INTERPRETED AS NO, 3-0f 0,f BLANK CHARACTERS OR A RETURN CARRIAGE ARE INTERPRETED AS A ZERO VALUE ORf WHEN APPLICABLEf ARE REPLACED BY DEFAULT VALUES. START ? ? yes ••d-'d-o-'O ENTER COMMENTS LINES, (MAXIMUM OF TERMINATE WITH END AT BEGINNING OF A NEW LINE, base test file (appendix C) exasple for chapter 3 toner Sroundind plus one dround potential control rind ENTER RUN IDENTIFICATION, (4 DIGITS) ? ex2 SYSTEM OF UNITS -METRIC SYSTEM ? ? no -BRITISH SYSTEM ? ? ses ? yes IS YOUR TOWER (OR POLE) GROUNDING SYSTEM SYMMETRICAL ? VIDEO TERMINAL OR HARD-COPY TERMINAL (PRINTER) FOR OUTPUT ? VIDEO ? ? yes ENTER SOIL DATA ENTER TOP AND BOTTOM LAYER RESISTIVITIES (OHMS-METER) AND TOP LAYER HEIGTH, (METERS OR FEET) IF UNIFORM SOILf ENTER TOP LAYER RESISTIVITY ONLY, EXAMPLE !R01fR02fH OR RO ? 383»5f147»66f8,405 ? ENTER FAULT CURRENT IN TOWER GROUNDING, 1000 C-6 COHFIIMATION OF TOWER GROUNDING ••'d —d**3'*«9 A- HORIZONTAL CONDUCTORS, <HAX. OF 16) - ENTER THE (X»Y) COORDINATES OF CONDUCTOR EXTREMITIES (ORIGIN AND END) THEN THE RADIUS FOLLOWED BY THE BURIAL DEPTH OF THE CONDUCTOR, - ENTER EACH NEW CONDUCTOR ON A SEPARATE LINE, - EXAMPLE i XOfYOfXEfYEiRABfDEPTH, - ALL VALUES ARE IN METERS OR FEET, - TERMINATE BY TYPING END AT THE BEGINNING OF ft NEW LINE IS.SjlZ.SflZarO.SfO.OAfS.A 14.1>12,9,12,9,14,1,0,04*5.4 14.3.13.1.13.1.14.3.0. 04.5.4 14.5.13.3.13.3.14.5.0. 04.5.4 i 0.0,18,5,18.5,18.5,0.04,0.75 0.18.5.18.5.0.04.0.75 ? 18.5.0. ? end B- VERTICAL CONDUCTORS, (MAX, OF 8) - ENTER THE (X,Y> COORDINATE OF THE CONDUCTOR, THEN THE RADIUS, THE LENGTH AND TOP END DEPTH OF THE CONDUCTOR. - ENTER EACH NEW CONDUCTOR ON A SEPARATE LINE, - EXAMPLE I X,Y,RAD,LENGTH,DEPTH - ALL VALUES ARE IN METERS OR FEET - TERMINATE BY TYPING END AT THE BEGINNING OF A NEW LINE, ? 14.,14,,0,04,5.36,0,04 ? end POTENTIAL PROFILES AT THE SURFACE OF SOIL,(MAX,OF 10) - ENTER THE <X,Y> COORDINATES OF PROFILE EXTREMITIES (ORIGIN AND END) FOLLOWED BY THE SPACING BETWEEN TWO CONSECUTIVE POINTS. - EXAMPLE I XO,YO,XE,YE,SPACING - START EACH PROFILE SPECIFICATION ON A NEW LINE. - ALL VALUES ARE IN METERS OR FEET - TERMINATE BY TYPING END AT BEGINNING OF A NEW LINE, ? 14.0,0.0,14.0,30.0,0.5 ? end END OF INPUT DATA FOR PROGRAM GTOWER, C-7 GTOWER PAGE 5 1 RUN ! EX2 tttmmtmmmmmtmmm IBASE TEST FILE (APPENDIX C) LE_F0R_CHAPTER_3£we cohhents mtmmmmmnnmtmum % POTENTIAL CONTROL RING cohhents ! ntmmmmmntnnmmtm RUN IDENTIFICATION ( EX2 UNITS SYSTEH 5 BRITISH SYHHETRICAL CODE 5 YES POTENTIAL OPTION 5 YES GTOWER PAGE { 2 RUN 5 EX2 SOIL DATA I--------* - TOP LAYER RESISTIVITY BOTTOH LAYER RESISTIVITY TOP LAYER HEIGTH REFLECTION FACTOR t 5 J t 383,500 OHHS-HETER 147.660 OHHS-HETER 8.405 FEET -.444 FAULT CURRENT I------------ * - IF i 1000,000 AHPERES GTOWER PAGE i 3 RUN : EX2 CONDUCTORS DATA - HORIZONTAL CONDUCTORS t---------- 1 YO FEET *- - - - - $ XE FEET ---- S — YE FEET *- - - s 13.5000 12.3000 12,3000 !?:» 12.9000 13.1000 13,3000 18,5000 18,5000 13,5000 i!:W 14,1000 14,3000 14.5000 18.5000 18,5000 XO FEET 14.1000 14.3000 14.5000 0.0000 18.5000 KRB 12,9000 13,1000 13.3000 18.5000 0.0000 RADIUS FEET DEPTH FEET t---------- 1 *---------- ? 0400 5.4000 m 5,4000 5.4000 5.4000 ,7500 .7500 a 0400 0400 0400 0400 0400 C-8 - VERTICAL CONDUCTORS 1------ 1 Y FEET *---------- « RADIUS FEET *---------- 1 LENGTH FEET *---------- S DEPTH FEET 1---------- 1 14.0000 14.0000 ,0400 5.3600 ,0400 X FEET GTOWER PAGE I 4 RUN \ EX2 PROFILES DATA *-------------- $ --------END OF PROFILES------YE XE FEET FEET t---------- $ t---------- 1 ------ ORIGIN OF PROFILES----YO XO FEET FEET 1---------- 1 t---------- 1 14.0000 14,0000 0,0000 30,0000 SPACING BETWEEN POINTS IF) t---------- 1 ,5000 END OF INPUT DATA AVERAGE NUMBER OF SERIES TERMS USEB= 5 GTOWER PAGE i RUN t EX2 * * I COMPUTER RESULT % t * GROUND POTENTIAL RISE GPR i 1216243 VOLTS RESISTANCE OF TOWER GROUNDING *---------------------------------------- —$ RESISTANCE X 124621 OHMS AVERAGE NUMBER OF SERIES TERMS USED= 5 C-9 5 BTQHEK PAGE J 6 RUN I EX2 POTENTIAL PROFILE 1 1 t----------- — -------------1 POINT NUMBER 1 2 3 4 5 6 ? 8 9 10 11 12 13 14 15 U 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 # 38 39 40 41 42 43 44 45 46 47 48 49 50 51 X FEET 14.000 14.000 14.000 14.000 14.000 14,000 14,000 14,000 14,000 14.000 14,000 14.000 14,000 14.000 14.000 14,000 14.000 14.000 14,000 14,000 Y FEET -- * 0.000 .500 1,000 1.500 2,000 2.500 3.000 3.500 4,000 4.500 5.000 5,500 6,000 6.500 7,000 7.500 8.000 8.500 9.000 9.500 14,000 14.000 14,000 14.000 14.000 14,000 14.000 14.000 14,000 14.000 14,000 14,000 14,000 14,000 14,000 11* 10.000 14.000 14,000 14,000 14,000 14,000 14.000 14.000 14.000 14.000 14,000 14,000 14,000 14,000 14.000 18.500 19.000 19,500 20,000 20.500 21.000 21,500 22.000 22,500 23.000 23,500 24,000 24.500 25,000 10.500 11.000 11.500 12.000 12,500 13.000 13,500 14.000 14.500 15.000 15.500 16,000 16,500 17,000 11* POTENTIAL VOLTS PERCENT OF GPR --- ------ 4 4 6189,711 6191,512 6196.936 6206.052 6218,978 6235.941 6257.046 6282.640 6313.074 6348.772 6390.232 6438.031 6492,832 6555.393 6626.771 6707.566 6798.145 6901.764 7019.012 7150,655 7301,250 7472.154 7668,011 7896,352 8169.292 8510,670 8959,655 9673.174 11620,592 9865.886 9356,276 9131,764 9055.782 9105,393 9275,290 50.893 50,908 50.953 51.028 51,134 51.273 51,447 51.657 51.908 52,201 52,542 52.935 53.386 53.900 54.487 55,151 55,896 56,748 57,712 58,794 60.033 61,438 63.048 64,926 67.170 69,977 73.668 79,535 95.547 81,120 76.930 75,084 74,459 74.867 76.264 Ml m 10119,229 9556.326 8721,292 7981,065 83,203 78,574 71.709 65.622 60,532 56.256 52,590 49,414 46,630 44.167 41,969 39,996 38,240 36,622 7362,027 6841,881 6396.065 6009.817 5671.253 5371.629 5104.349 4864,312 4650.855 4454,009 C-10 GTOWER PAGE 5 RUN { EX2 POINT NUMBER --------- * X FEET t---------- $ Y FEET *----- -—| POTENTIAL MOLTS 1---------- * I--------- 1 14.000 14.000 14.000 25.500 26,000 26.500 i« 2:IH 14.000 14.000 14.000 14.000 14.000 23.000 28.500 29,000 29.500 30,000 4274,499 4110,144 3959.120 3819,890 3691.145 3571.763 3460.775 3357,342 3260.730 3170,295 35,146 33,795 32,553 31,408 30.350 29,368 28.455 27,605 26.811 26,067 C-11 PERCENT OF GPR 7 GTOUER PAGE RUN 5 EX2 , 0, I__ 1300. 3900, 2600. 5200, 6500. 7800. .J______ i______ I______ I______ I______ I___ t 8 POTENTIAL (MOLTS) 9100. 10400, 11700, 13000. ..I______ I______ I__- _I + 0 0000-1 I I I I + + * + + + I 1 I t + 6.0000-1 + i s :o a 3 T i m o a e ^ —««««-* es I I * * f 1 I I I 1 + 4 4 12,0000-1 + I I I I I I 1 I I 4 « rs ir-fr-t-n o s s -o 18,0000-1 1 I 1 1 I I 4 4 + I 24,0000-1 ( F ) I I 1 I I I I 1 I 4 4 4 4 4 4 4 4 4 4 * 4 4 4 , GRID POTENTIAL RISE 4 EARTH SURFACE PROFILE # 1 EX2 PROFILE BASED ON 1000,00 AHPERES TONER CURRENT C-12 « GTfflO PAGE 5 RUN I EX2 L 0, , 10 , 20 30. 40, 50. T 0,0000-1 I I I I I I I I I 6.0000-1 I I ■T + * t + i 1 1 I I I I I I I I I I I I I I 18.0000-1 I I I I I I I I I 12.0000-1 24,0000-1 1 I I I I I I I I , GRID POTENTIAL RISE + EARTH SURFACE PROFILE # 1 EX2 NQRHALIZED PROFILE BASED ON 100 VOLTS GPR END OF GTtttER mmmtm C-13 60. POTENTIAL (VOLTS) ^70, ^80, ^90, 9 100, ( APPENDIX D COMPUTER PROGRAM PATHS FUNCTION PATHS is used to compute currents in the ground wire and in the structures of a single or double circuit transmission line when a fault occurs at any structure on the line. Each line extremity is connected to a terminal representing the equivalent network at the end of the line. Single, double and three phase-to-ground faults are permitted. REFERENCES Information on this program is given in Chapter 2. PATHS is based on the generalized double-elimination method described in Chapter 6. A typical application of PATHS is described in Chapter 3. Another example is analysed in Chapter 10. GENERAL PATHS was developed for use in an interactive mode. In this mode, the user is prompted to provide the required input data and answers in free-format. A typical interactive session is shown in Page D-10. The program can also be readily operated in batch mode. In this mode, the user must organize the data in a file prior to running the program. The batch mode is particularly advantageous when running several cases with small differences in the input data. The following guidelines and the interactive session shown in Page D-10 provide the necessary instructions to build an input data file for the program PATHS. GUIDELINES All values entered on a data line must be separated by a comma. Blanks are interpreted as a YES, zero or default answer, depending on the data expected by the program. PATHS accepts up to 28 different types of data lines. Occasionally, several data lines of the same type are accepted by the program. In this case, the user must enter a line consisting of the world END as the last entry to instruct the program that the series of data lines of the current type has ended. A typical input data file is shown in Page D-9. This file can be prepared based on the following guidelines: D-l Data Line 1(a) : Start This line must necessarily be a YES to initiate execution of the program. A NO will stop the program. Data Line 1(b) : Detailed Explanations Enter YES if you need detailed explanations (during interactive sessions) and NO otherwise. Data Line 2 : Comment Enter any alphanumeric character to provide comments and general references. These lines are not processed by the program but simply reprinted at the top of the output printout. Up to four lines of comments may be entered. If less than four comment lines are entered, the comments must be terminated with a line consisting of the word END. If the character is among the characters entered in a comment line, all the information to the right of this character is ignored. This is a useful feature (for both interactive and batch modes) which allows the user to include additional comments in the input data which will not appear in the printout. Data Line 3 : Run Identification Enter a four digit identification code which will be printed at the top right-hand side of each page of printout. Data Line 4 : System of Units If the metric system is to be used, provide a data line beginning with a YES. If the British system of units is requested, two data lines are needed; line 1 is a NO and line 2 is a YES. Data line 5 : Detailed Printout A YES in this line is a request for computation of the currents in every structure and ground wire span of the transmission line. If a NO is entered in this line, data lines 6 to 8 should be skipped because the optional plots produced by PATHS only display the detailed computation results in graphical form. Data Line 6 : Plots of Structure Currents Enter a YES if plots of the currentin a structure, as a function of the structure (tower) number, are required. A NO must be entered if the plots are not required. Data Line 7 : Plots of Span Currents Enter a YES if a plot of ground wire span currents on each side of the faulted structure as a function of the span number, is required. Enter a NO if this is not required. D-2 Data Line 8 : Type of Terminal Used Skip this line if a NO was entered in lines 6 and 7. If the terminal is a video screen, only one character at a time can be displayed in a specific location. This is not the case of printers and hard-copy terminals which allow overprinting. The program uses two different routines to produce plots, depending on the type of terminal. The hard-copy terminal routine produce the most accurate plots. However, the other routine can be used on a video or a hard-copy terminal. This line should be a YES for a hard-copy terminal. Two lines, a NO followed by a YES, are needed to select the video terminal. Data Line 9 : Zero Impedances True zero impedance values cannot exist in reality. In order to avoid computation errors due to zero impedance loops, PATHS modifies zero resistance and inductance values to 0.001 ohms. This does not apply to mutual impedances. If zero impedance values are required, this line should be a YES, otherwise NO. Data Line 10 : Line Transposition Enter YES if transmission line conductors are transposed, otherwise NO. When the transposed transmission line is so specified, PATHS assumes that all mutual impedances between the phase conductors of a circuit are equal and that the mutual impedances between the equivalent ground wire and the phase conductors of a circuit are also equal. Data Line 11 : Single or Double Circuit Enter a YES for a single circuit transmission line. Enter NO and then YES in the next line if the problem involves a double circuit transmission line. Data Line 12 : Line Voltage This is a pair of values separated by a comma. The first value represents the real part of the line voltage of phase A of circuit 1 (in kV). The second value is the imaginary part of this voltage (in kV). Generally, it is easier to take this voltage vector as the reference vector (zero imaginary part). The voltages of the other phases are automatically generated by PATHS. Data Line 13(a) : Nature of Left Terminal Enter a YES if the left terminal is a power source (contributes a current into the fault) with respect to circuit No. 1, otherwise NO. D-3 Data Line 14(a) : Left Terminal Data Enter the following data: NAME, RG, IG, REQ1, IEQ1, REQ2, IEQ2 where NAME is the name of the terminal. RG+jIG is the ground impedance of the grounding system of the terminal (in ohms). REQl+jlEQl is the equivalent system impedance of the terminal for circuit No. 1 (in ohms). REQ2+jIEQ2 is the equivalent system impedance of the terminal for circuit No. 2 (in ohms). Data Line 15(a) : Number of Structures on Left Side Enter the number of structures between the faulted structure, No. 0, and the left terminal. This value must be greater than zero. Data Line 13(b) : Nature of Right Terminal Similar to 13(a) Data Line 14(b) : Right Terminal Data Similar to 14(a) Data Line 15(b) : Number of Structures on Right Side Similar to 15(a) Data Line 16(1) : Phase Impedance of Circuit 1 Enter the resistance and reactance of the phase conductors of circuit No. 1 (in ohms/km or in ohms/mile). This is the average value of the phase impedances computed by a line parameter program. If the transmission line is not transposed, skip the following data lines 17, 18 and 19. D-4 Data Line 17(1) s Phase-to-Phase Mutual Impedance This is the average value of mutual impedance between phase conductors of circuit No. 1 (resistance and reactance in ohms/km or ohms/mile). Data Line 18(1) : Ground Wire-to-Phase Mutual Impedance This is the average value of mutual impedance between the ground wire and the phase conductors of circuit No. 1 (in ohms/km or ohms/mile). Data Line 16(2) : Phase Impedance of Circuit 2 Similar to 16(1) but for circuit No. 2 (if any). Data Line 17(2) : Phase-to-Phase Mutual Impedance Similar to 17(1). Data Line 18(2) : Ground Wire-to-Phase Mutual Impedance Similar to 18(1) Data Line 19 : Circuit-to-Circuit Mutual Impedance This is the average value of the mutual impedance between the phase conductors of circuit No. 1 and circuit No. 2 (in ohms/km or ohms/mile). This. data line must be entered only in the case of a double circuit line. Skip the following data line (20) if the transmission line was specified to be transposed. Data Line 20 : Mutual Impedances of Untransposed Lines Because the transmission line is untransposed, the mutual impedance between each pair of conductors must be specified. The unspecified values are assumed to be zero. The mutual impedance value between two conductors is specified as follows: ID1, ID2, RMU, IMU where ID1 and ID2 identify the first and the second conductor respectively. Al, B1 and Cl designate phases A, B and C of circuit No. 1 and GW designates the ground wire. RMU+jXMU is the mutual impedance (in ohms/km or ohms/mile). When the mutual impedance of all conductor pairs have been specified, an END line must be entered to instruct the program to move to the next type of data. Data Line 21 : Ground Wire Self Impedance Enter the resistance and inductance of the impedance of the ground wires or lumped equivalent ground wire in ohms/km or ohms/mile. D-5 Data Line 22 : Faulted Structure Impedance Enter the resistive structure (in ohms). and inductive components of ground impedance of the faulted Data Line 23 : Fault PATH Impedances The program PATHS assumes that there is a connection between the phase conductors, ground wire and the "faulted" tower structure. These connections define the fault paths. By default, PATHS assumes an infinite impedance path (open connection) between the phase conductors and the structure and a zero impedance path (solid metallic connection) between the ground wire and the structure. This corresponds to a no fault condition. The user may simulate any type of fault by specifying impedance values other than the default values. It is only necessary to define those values which are different from the default values. For example a ground fault on phase A of circuit 1 is created by entering the following data line: Al, 0.0, 0.0 This implies that the fault impedance path between phase conductor Al and the structure is O.O+jO.O ohms. The general format of the data line is: ID, RCON, ICON where ID is the conductor identification as defined in data line 20. RCON+jICON is the fault path impedance (in ohms). A line beginning with END must be entered after all fault path impedances have been defined. Data Line 24(a) : Distances Between Left Side Structures Enter a YES if the structures on the left side of the faulted structure are to be considered egually spaced. In this case, the average spacing between two consecutive structures is specified in the next data line (type 25) and the ground impedances of the structures are specified by data lines of the type 26. If a NO is entered, specifying that the distance between two adjacent structures is variable along the left section of the line, then the distance between a structure and the adjacent structure to the left and the structure ground impedance must be defined by data lines of the type 27. Consequently, in the case of a YES, data line 27 must be skipped and for a NO, both data lines 25 and 26 must be ignored. Data Line 25(a) : Average Left Side Structures Spacing Enter the average spacing between two adjacent structures on the left portion of the transmission line (in m or ft). D-6 Data Line 26(a) : Left Side Structures Ground Impedances The general format of this line is: I, 3, RTOWER, ITOWER where I and 3 represent the group of structures bearing the numbers I and 3, all having the following ground impedance: RTOWER+jITOWER is the ground impedance of structures No. I to 3 (in ohms). An END line must be entered once all structures have been specified. Only one structure is specified when 1=3. It is possible to overwrite the specification of a structure by entering a new data line followed by a data line beginning with a NO to confirm that overwriting is not accidental. For example, if 1=5 and 3=21 in one data line and later a line with 1=17 and 3=18 is entered, then structures 17 and 18 will have the last specified values only if a data line with NO is entered immediately after the new structure specification. Data Line 27(a) : Left Side Structures Data This data line is used when the distances between structures are not equal. The general format of this line is: I, 3, SPA, RTOWER, ITOWER where I and 3 represent the group of structures bearing the numbers I and 3. These structures are all separated by the following distance and have the following ground impedance. SPA is the distance between two consecutive structures (in m or ft). RTOWER+jITOWER is the ground impedance of structures No. I to 3 (in ohms). A line beginning with END must be entered to instruct the program that all structures have been specified. The procedure for overwriting previously specified values is also applicable (see data line 26). Data Line 28(a) : Left Side last Structure Spacing This is the distance between the left terminal and the closest transmission line structure (in m or ft). This structure is the highest numbered structure of those on the left of the faulted structure (structure No. 0). D-7 Data Line 24(b) ; Distances Between Right Side Structures Similar to data line 24(a). Data Line 25(b) : Average Right Side Structure Spacing Similar to data line 25(a). Data Line 26(b) : Right Side Structure Ground Impedances Similar to data line 26(a). Data Line 27(b) : Right Side Structure Data Similar to data line 27(a). Data Line 28(b) : Right Side Last Structure Spacing Similar to data line 28(a). This is the last data entry of the program PATHS. D-8 INPUT DATA FILE FOR BATCH MODE PROCESSING Y Y BASE TEST FILE <APPENDIX D> EXAMPLE FOR CHAPTER 3 SINGLE CIRCUIT NO RING EX I N Y Y Y Y N Y Y Y Y 345 Y TlrO.01*0.0**99*14.85 14 Y T2»0.01*0.0*2.64*39.59 85 0.141*1.169 0.091*0.579 0.090*0.536 0.808*1.234 31.67 A1*0.0*0.0 END Y 800. 1*1*27*O 2*3*24 *O 4*5? 22 * O 6*14*20*0 END 800 ♦ Y 800. 1*1*28*0 2*3*27*0 4*7*24*0 8*10*22*0 11*85*20*0 END 800 . D-9 TYPICAL INTERACTIVE DIALOGUE AND COMPUTER PROGRAM PRINTOUT PR08RAH PATHS %%%%%%%%%%%%% THIS PROGRAM CALCULATES THE DISTRIBUTION OF FAULT CURRENT IN A TRANSMISSION LINE SUBJECT TO A FAULT ON A TOWER. NOTES : 1-RESPOND TO QUESTIONS BY YES OR NO OR. WHEN APPROPRIATE. BY TYPING THE REQUIRED DATA, 2-Yf YE. YES. BLANK CHARACTERS AND RETURN CARRIAGE ARE INTERPRETED AS YES. OTHER CHARACTERS ARE INTERPRETED AS NO, 3-0. 0., BLANK CHARACTERS OR A RETURN CARRIAGE ARE INTERPRETED AS A ZERO VALUE OR. WHEN APPLICABLE. ARE REPLACED BY DEFAULT VALUES, START ? yes DO YOU WANT A PRINTOUT OF THE GUIDELINES BEFORE STARTING THE CONVERSATIONAL SESSION ? ? yes GENERAL GUIDELINES tttttttttttttttttt 1- YOU CAN MODEL A SINGLE OR DOUBLE CIRCUIT TRANSMISSION LINE WITH NO OR ONE OVERHEAD GROUND-WIRE, IF THERE ARE SEVERAL GROUND-WIRES, FIND AN EQUIVALENT FICTITIOUS 0H6W CONDUCTOR (YOU CAN USE BUNDLE REDUCTION TECHNIQUES OR EQUIVALENT (SEE COMPUTER PROGRAM LINPA), 2- ALL IMPEDANCES AND VOLTAGES MUST BE ENTERED IN CARTESIAN FORM, I.E., FIRST ENTER REAL PART THEN IMAGINARY PART OF NUMBER. EXAMPLE ’, RNUMBER, INUMBER 3- IF THE PHASE CONDUCTORS OF ONE CIRCUIT ARE NOT TRANSPOSED YOU MUST ENTER THE MUTUAL IMPEDANCE BETWEEN ALL PHASE AND GROUND WIRE CONDUCTORS OF ALL CIRCUITS (EXCEPT WHEN ZERO), HOWEVER WHEN THE PHASE CONDUCTORS OF ALL CIRCUITS ARE DECLARED TRANSPOSED, YOU MUST ENTER ONE AVERAGE VALUE FOR THE MUTUAL IMPEDANCE BETWEEN i A- ALL B- THE C~ THE (*> TOTAL PHASES OF A CIRCUIT*, AND PHASES OF CIRCUITS 1 I 2 (DOUBLE CIRCUIT), AND PHASE CONDUCTORS AND THE GROUND-WIRE FOR EACH CIRCUIT* OF TWO VALUES FOR A DOUBLE CIRCUIT LINE TO ENABLE THE TRANSPOSITION OPTION, ANSWER IYESI TO THE APPROPRIATE QUESTION (SEE BELOW) 4- WHEN THE PHASE CONDUCTORS ARE NOT TRANSPOSED THE MUTUAL IMPEDANCES ARE SPECIFIED USING THE FOLLOWING CONDUCTOR IDENTIFICATION CODE ! Al- PHASE A OF CIRCUIT LINE VOLTAGE ANGLE Bl- PHASE B OF CIRCUIT LINE VOLTAGE ANGLE Cl- PHASE C OF CIRCUIT LINE VOLTAGE ANGLE A2- PHASE A OF CIRCUIT VOLTAGE ANGLE jo_ LINE t 1 = 1 = 1 = 2 = 0 +120 SHIFT +240 SHIFT 0 GW- EQUIVALENT SINGLE OVERHEAD GROUND-WIRE D-10 5- ft SECTION. IS DEFINED AS THE PORTION OF TRANSMISSION LINE BETUEEN TWO GROUNDED POINTS OF THE GROUND-WIRE AND THE GROUNDING SYSTEM (USUALLY A TOWER GROUND) AT THE SECTION'S EXTREMITY WHICH IS OPPOSITE TO THE FAULT LOCATION. NORMALLY THESE POINTS CORRESPOND TO A TRANSMISSION LINE TOWER WICH IS CONNECTED TO THE GROUND WIRE. 6- YOU CAN SPECIFY IN ONE DATA LINE ONE OR SEVERAL CONSECUTIVE SECTIONS WITH IDENTICAL CHARACTERISTICS. YOU ARE ALLOWED TO ENTER A MAXIMUM OF 200 SECTION DATA LINES. NO LIMITS ARE ASSIGNED TO THE NUMBER OF SECTIONS PROVIDED HOWEVER THAT THE NUMBER OF SECTION DATA LINES IS NOT EXCEEDED 7- ALL UNSPECIFIED MUTUAL IMPEDANCE VALUES ARE ASSUMED ZERO ALL UNSPECIFIED SECTION DISTANCES ARE ASSUMED EQUAL TO THE LAST SPECIFIED SECTION DISTANCE, ALL UNSPECIFIED TOWER IMPEDANCES ARE ASSUMED AS 99999.+J0. UNLESS THE ZERO IMPEDANCE ACCEPTANCE OPTION IS ACTIVATED ANY ZERO RESISTANCE OR INDUCTANCE SPECIFIED BY THE USER WILL BE TAKEN AS 0.001 8- TO SIMULATE PHASE TO TOWER FAULT» SPECIFY AN APPROPRIATE LOW IMPEDANCE CONNECTION BETWEEN THE FAULTED PHASE(S) AND THE TOWER STRUCTURE, IF NOTHING IS SPECIFIED FOR A GIVEN PHASE CONDUCTOR THEN THE PROGRAM ASSUMES NO CONNECTION BETWEEN THIS PHASE AND TOWER STRUCTURE BY DEFAULT THE EQUIVALENT GROUND WIRE IS ASSUMED TO HAVE A PERFECT CONNECTION TO THE TOWER STRUCTURE, HOWEVER YOU CAN SIMULATE A BAD OR OPEN CONNECTION (GW-TOWER) BY ENTERING A SUITABLE IMPEDANCE VALUE FOR THIS GROUND WIRE TO TOWER CONNECTION ENTER COMMENT LINES. (MAXIMUM OF 4) TERMINATE WITH iENDt AT BEGINNING OF A NEW LINE, base test file (appendix D) example for chapter 3 single circuit no around potential control rind ENTER RUN IDENTIFICATION. (4 DIGITS) ? exl SYSTEM OF UNITS - METRIC ? ? no - BRITISH ? ? yes DETAILED PRINTOUT OPTION. CURRENT IN EACH TOWER WILL BE CALCULATED MS PRINTED ? ? yes DO YOU WANT PLOTS OF I 1- FAULT CURRENT DISTRIBUTION IN TOWERS ? ? yes 2- FAULT CURRENT DISTRIBUTION IN OVERHEAD GROUND WIRE ? ? yes PRINTOUT ROUTED TOt - A PRINTER OR A HARD-COPY TERMINAL? ? no D-11 - ft VIDEO TERHINftl <NO OVERWRITE CAPABILITY) ? ? aes ARE YOU PLAHHIHG TO ENTER ZERO IHPEDANCE VALUES ? ? aes ARE YOUR PHASE CONDUCTTOS TRANSPOSED ? ? aes SINGLE OR DOUBLE CIRCUIT -SINGLE ? ? aes. ENTER THE LINE VOLTAGE IN KV (PHASE TO PHASE). EXAKPLE S RVOLTfIVQLT. ? 345 DATA FOR LEFT-SIDE OF FAULTED TOWER EACH CIRCUIT OF THE TERMINAL REPRESENTS A SOURCE OR A LOAD. IS THE TERMINAL A SOURCE WITH RESPECT TO CIRCUIT NUMBER ONE ? ? aes TERMINAL DATA ENTER THE NAME, THE GROUNDING SYSTEM IMPEDANCE (OHMS) AND THE EQUIVALENT TERMINAL IMPEDANCE (OHMS) OF CIRCUIT NUMBER ONE FOLLOWED (WHEN APPLICABLE) BY THE TERMINAL IMPEDANCE (OHMS) OF CIRCUIT NUMBER TWO. EXAMPLE! NAME'RGROUNDjIGRQUNDjREQII>IEQtliREQf2rIEQ*2 ? TliO.OlfO.Oi.99*14,85 ? 14 ENTER THE NUMBER OF GROUNDED TOWERS ON THIS SIDE OF FAULTED TOWER, NOTE ! TERMINAL IS NOT CONSIDERED AS A TOWER DATA FOR RIGHT-SIDE OF FAULTED TOWER EACH CIRCUIT OF THE TERMINAL REPRESENTS A SOURCE OR A LOAD, IS THE TERMINAL A SOURCE WITH RESPECT TO CIRCUIT NUMBER ONE ? ? aes TERMINAL DATA ENTER THE NAME. THE GROUNDING SYSTEM IMPEDANCE (OHMS) AND THE EQUIVALENT TERMINAL IMPEDANCE (OHMS) OF CIRCUIT NUMBER ONE FOLLOWED (WHEN APPLICABLE) BY THE TERMINAL IMPEBANCE (OHMS) OF CIRCUIT NUMBER TWO, EXAMPLEt NAME»RGRQUNB»IGROUND.REQHrIEQtl.REQ*2»IEQI2 ? T2.0,01,0,0»2.64.39.5? ENTER THE NUMBER OF GROUNDED TOWERS ON THIS SIDE OF FAULTED TOWER. NOTE 1 TERMINAL IS NOT CONSIDERED AS A TOWER CONDUCTOR IMPEDANCES ENTER THE PHASE IMPEDANCE OF CIRCUIT 1 IN OHMS/MILE EXAMPLE* RPHASE.IPHASE ? 0,141.1,169 ENTER PHASE TO PHASE MUTUAL IMPEDANCE FOR CIRCUIT 1 OHHS/HILE EXAMPLE! RMUTUPHASE.IMUTUPHASE ? 0,091.0.57? 0-12 ENTER PHASE TO GROUND HIRE HUTUAL IKPEDANCE Fffi CIRCUIT 1 OHHS/HILE EXAHPLE? RMUTU6R0UND»IMUTUGROIWD ? 0.090»0.536 ENTER GROUND HIRE IHPEDANCE IN OHHS/HILE EXAHPLE: RGHfIGU. f 0.808)1.234 ENTER THE FAULTED TWER GROUNDING IHPEDANCE (OHHS) EXAHPLE 1 RFAULT) FAULT ? 31.67 PROVIDE CONDUCTOR TO TOMER STRUCTURE CONNECTION (FAULT PATH) IHPEDANCE (WHS) UHEN DIFFERENT FROH DEFALT VAUES I.E.) PHASE CONNECTION ! OPEN GROUND HIRE CONNECTION ! SOLID EXAMPLE *, B2fRCON» ICON TERMINATE WITH IENDt AT THE BEGINNING OF A NEH LINE. AlfO.OfO.O ? end TOMER IHPEDANCE LEFT SIDE OF FAULTED TOMER ARE ALL SPACIN6S BETWEEN TOWERS EQUAL ? ? yes ENTER THIS SPACING.(METERS OR FEET) ? 800 ? ? ? ? ? ENTER THE TOMER IMPEDANCE (OHMS) EXAHPLE1 IrJ)RTOHER)ITOUER TERMINATE WITH END AT THE BEGINNING OF A NEH LINE lri)27)0 2)3)24)0 4)5)22)0 6)14)20)0 end WHAT IS THE DISTANCE BETWEEN THE LAST TWER AND TERMINAL STATION ? (H OR FT) ? 800 RIGHT SIDE OF FAULTED TOMER ARE ALL SPACINGS BETWEEN TOWERS EQUAL ? ? ses ENTER THIS SPACING.(METERS OR FEET) ? 800 ENTER THE TWER IMPEDANCE (OHMS) example: I)J)RtoheR)Itower TERMINATE WITH END AT THE BEGINNING OF A NEH LINE ? 1)1)28)0 ? 2)3)27)0 ? 4)7)24)0 ? 8)10)22.0 ? 11)85.20)0 ? end WHAT IS THE DISTANCE BETWEEN THE LAST TOMER AND TERMINAL STATION ? (M OR FT) ? 800 D-13 PATHS PAGE *. 1 RUN 5 EX1 INPUT DATA mmmt t * t t t BASE TEST FILE (APPENDIX D) EXAMPLE FOR CHAPTER 3 SINGLE CIRCUIT m RING % RUN IDENTIFICATION ! EX1 UNITS SYSTEM t BRITISH TYPE OF CIRCUIT 5 SINGLE PRINTOUT ON TERMINAL OPTIONS SELECTED ZERO IMPEDANCE ACCEPTANCE t YES TRANSPOSED PHASE CONDUCTORS 5 YES DETAILED OUTPUT I YES PLOT OF CURRENT DISTRIBUTION IN TOWERS S YES PLOT OF CURRENT DISTRIBUTION IN GROUND WIRE I YES PATHS PAGE 1 2 RUN S EX1 CIRCUIT DESCRIPTION LEFT TERMINAL (NO.l) DATA GROUNDING SYSTEM IMPEDANCE 5 CIRCUIT NUMBER ONE EQUIVALENT SOURCE IMPEDANCE 1 SOURCE TERMINAL. VOLTAGE I .990 +J 345.000 +J .010 +J 0.000 OHMS 14.850 OHMS 0.000 KILOVOLTS RIGHT TERMINAL (N0.2) DATA NAME S T2 GROUNDING SYSTEM IMPEDANCE 5 .010 « CIRCUIT NUMBER ONE EQUIVALENT SOURCE IMPEDANCE 5 SOURCE TERMINAL. VOLTAGE ! 2.640 +J 345.000 +J 0.000 OHMS 39.590 OHMS 0.000 KILOVOLTS D-14 SELF 1HPEMNCE CIRCUIT 1 {PHASE CONDUCTORS) S OVERHEAD EQUIVALENT GROUND HIRE l ,141 +J ,808 +J 1,16? OHHS/HILE 1,234 OHHS/HILE CONDUCTOR TO FAULTED STUCTURE CONNECTION IMPEDANCE PHASE Al TO TONER STRUCTURE PHASE Bl TO TONER STRUCTURE PHASE Cl TO TOWER STRUCTURE GROUND WIRE TO TOWER STRUCTURE ! 5 0,000 +J ------- OPEN ------- OPEN 0.000 +J l 1 0*000 — 0*000 PATHS PAGE I 3 RUN : EX1 HUTUAL IMPEDANCES CONDUCTOR OF CIRCUIT <T0> CONDUCTOR OF CIRCUIT Bl Bl 1 1 1 Al 1 GN Bl 1 Cl 1 Al Al IMPEDANCE OHHS/HILE .091 « .579 ,091 +J .579 .091 +J ,579 .090 FJ ,536 GW 1 1 1 0 0 .090 4J .536 GN 0 ,090 +J .536 Cl Cl PATHS PAGE I 4 RUN i EX1 GROUND IMPEDANCE OF TONERS FWILTED TONER *, 31,670 +J 0.000 OHHS LEFT SIDE OF FAULTED TONER NUMBER OF GROUNDED TONERS S TOWER TO TONER NO. NO. 1 2 4 6 1 3 5 14 14 TOWER GROUND IHPEDANCE OHHS 27,000 24,000 22,000 20,000 DISTANCE FROM LAST TONER (NO. +J tJ +J W DISTANCE FROM PRECEDING TOWER FEET 0,000 0.000 0,000 0.000 800.000 800.000 800,000 800,000 14 ) TO TERMINAL T1 D-1 5 \ 800.000 FEET RIGHT SIDE OF FAULTED TWER NUNBER OF GROUNDED TOUERS 1 3 7 10 85 4 8 11 DISTANCE 85 TOWER GROUND IMPEDANCE OHMS TOWER TO TOWER NO. *N°' 1 2 i 28.000 27.000 24.000 22.000 20.000 +J « « +J +J DISTANCE FROM PRECEDING TOWER FEET ^ 0.000 0.000 0.000 0.000 0.000 800.000 800.000 800,000 800.000 800.000 FROH LAST TtSiER (NO. 85 ) TO TERMINAL T2 I 800.000 FEET END OF INPUT DATA PATHS PAGE i 5 RUN 5 EX1 COMPUTATION RESULTS mmmtmmm CURRENT DISTRIBUTION AT FAULTED TOWER NATURE OF CURRENT MODULE AMPERES ANGLE RADIANS ______ ____________ * * TOTAL FAULT t TOTAL EARTH t 15272.103 376.073 t t LEFT SIDE GROUND WIRE * RIGHT SIDE GROUND WIRE % % LEFT SIDE PHASE Al 8936,025 5993,235 RIGHT SIDE PHASE Al 11499,338 3773,411 % LEFT SIDE PHASE Bl 114.460 116.113 t -1.45 t -1,07 t $ 1,71 * 1.65 $ * -1,45 * -1,44 1 Jr * RIGHT SIDE PHASE Bl t I III ME 11 1 t------------- -------- --- ------ * FWILTED TOWER POTENTIAL RISE t -1.72 * 1.43 t * 1:8 f t ----------------------------------------- 1 11910.243 V.» < -1.07> RADIANS D-16 PATHS PAGE 5 6 RUN ! EX1 DETAILED PRINTOUT %%%%%%%%%%%%%%%%% CURRENT IN EACH SECTION AND TOWER LEFT SECTION NO. t- SIDE OF FAULTED TOWER SECTION CURRENT RODULE ANGLE AMPERES RADIANS TOWER NO. TOWER CURRENT MODULE ANGLE AMPERES RADIANS t t * % t t t t t t 1 2 3 4 5 6 7 8 9 t iO * 11 * 12 13 t 14 t % * 15 8936.025 8557,254 8169,789 7819.224 7475.273 7168.133 6869.129 6607.428 6381.656 6190.642 6033.404 5909.143 5817.228 5757.196 5728.742 1.71 1.70 1.68 1.67 1.66 1.65 1.63 1.63 1.62 1.61 1.61 1.60 1.60 1.60 1.60 1 2 3 4 5 6 7 8 9 10 11 12 13 14 400.390 407.153 366.369 357.612 317.858 308.139 268.733 t 1 2.04 t 2,00 t 1.97 t 1,94 * 1.91 t 1.88 * 1,86 * l.'HJ 160.195 126.380 93,359 60,919 28.856 1,80 1.79 1,78 1.77 1.76 t t t * * t - - - - -t t- D-17 PATHS PAGE 5 7 RUN I EX1 RIGHT SIDE OF FAULTED TOUER SECTION NO. *1 t t % t % t t % t t t t TOUER NO. TOUER CURRENT MODULE ANGLE AMPERES RADIANS ----- 1 — i 2 3 4 5 6 7 8 9 10 11 12 i 13 t 14 * 15 t 16 l 17 * 18 * 19 t 20 * 21 t 22 $ 1 $ * * l l 1 SECTION CURRENT ANGLE MODULE AMPERES RADIANS 23 24 25 26 27 28 29 30 5993.235 5631.547 5284.852 4964.605 4632.166 4325.692 4043.322 3783.372 3522.627 3283.152 3063.499 2842.257 2640.000 2455.466 2287,520 2135,140 1997.398 1873,443 1762,487 1663.783 1576.612 1500.267 1434.040 1377.217 1329.069 1288,853 1255,821 1229,226 1208,332 1192.428 1.65 1,62 1,60 1.58 1,55 1.53 1.51 1,50 1,48 1.47 1,46 1.45 1.45 1.45 1,45 1,46 1.47 1.49 1,50 1.53 1.55 1,58 1.61 1,64 1.68 1.71 1.74 1.78 1,81 1,84 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 390,626 371.612 340.549 350,699 320,815 293.265 267.864 266.670 243,242 221.759 222,268 202.525 184,535 168.143 153,206 139,595 127,193 115.892 105,594 96,211 87,661 79.870 72,771 66,302 60.408 55,037 50.143 45.683 41.620 D-18 2.03 1,98 1.94 1,89 1,84 1,80 1.75 1,70 1.65 1,60 t % » * t * * t * t » t 1,55 * 1,50 1,45 1.40 1.35 1.30 1.25 1,20 1.15 1.10 1,05 1.00 .95 .90 ,85 * t t t « * $ * * t * t * t .80 « ,75 1 ,70 * ,65 * PATHS PAGE t 8 RUN t EXi RIGHT SIDE OF FAULTED TOWER (CONTINUED) SECTION NO* i % t % t t 31 32 33 34 35 35 37 38 39 40 % 41 t t * % t t 42 43 44 45 46 47 t 48 t 49 * 50 t % 51 l 52 i % t * 53 54 55 56 t 57 t 58 59 60 61 * 62 t 63 t t t t 64 t #55 1180,839 1172,598 1167,758 1165,501 1165,379 1166,985 1169,960 1173,998 1178.837 1184.256 1190,072 1196.133 1202,313 1208.512 1214.648 1220,659 1226.494 1232.116 1237,498 1242.620 1247.472 1252.045 1256,340 1260,357 1264,101 1267.580 1270,802 1273,779 1276,521 1279,041 1281.351 1283.464 1285,393 1286,624 1288,224 1.87 1,90 1.93 1.95 1.98 2.00 2.01 2.03 2.05 2.06 2.07 2.08 2,09 2,10 2.11 2,11 2.12 2.12 2,12 2.13 2.13 2.13 2.13 2.14 2.14 2.14 2.14 2,14 2,14 2,14 2.14 2.14 2.14 2,14 2.14 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 58 59 60 61 62 63 64 37,917 33.776 31.591 28.812 26,254 23.916 21.785 19,844 18,075 16,463 14.995 13.656 12.437 11.325 10.313 9.390 8,549 7.783 7.085 6.449 5.870 5.343 4.863 4.426 4.029 3,668 3,340 23$5 .60 .56 ,50 .45 ,40 .35 .30 .25 ,20 ,15 .10 .05 -.01 -.06 -.11 -.16 -.21 -.26 -.31 -.36 -.41 -.47 -.52 -.57 -.63 -.68 -.74 2.531 2.311 2.114 1,937 1.597 1.617 -.91 -.97 -1.03 -1.09 -1.69 -1.15 D-19 t f t TOWER CURRENT MODULE ANGLE AMPERES RADIANS t * % % 1 * t t t * t t * * t % t * * t * * t t t $ t CJT-O t t TOUER NO, 03 t t * SECTION CURRENT ANGLE AMPERES RADIANS MODULE * * * t t * PATHS PAGE J f RUN 5 EXI RIGHT SIDE OF FAULTED TONER (CONTINUED) SECTION NO. if t t M 67 68 % 69 * 70 t t * 71 t t 72 73 * 74 t 75 76 % 77 % 78 t t t % % 79 80 81 82 83 % 84 t 85 * 86 * % SECTION CURRENT MODULE ANGLE AMPERES RADIANS 1289.672 1290.993 1292.199 1293.301 1294.310 1295.235 1296.088 1296.878 1297.615 1298.310 1298.972 1299.611 1300.237 1300.859 1301.488 1302,133 1302,805 1303,515 1304,272 1305,087 1305,973 2,14 2.14 2.14 2,14 2,14 2,14 2.14 2,14 2,13 2,13 2,13 2.13 2.13 2,13 2,13 2.13 2.13 2.13 2.13 2.13 2.13 TOUER CURRENT MODULE ANGLE AMPERES RADIANS TONER NO, 65 1.474 1.366 1.274 1.195 1.128 1.074 1.029 .994 ,968 .950 ,938 ,932 ,933 ,938 .949 .965 .986 1.013 1.047 1.088 1.137 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 —------- -------------------------- ----------------------- —---------- -- ------------------------ -1,19 -1,26 -1.33 -1,40 -1.47 -1.54 -1.60 -1,66 -1,71 -1.76 -1,79 -1.82 -1.84 -1.85 -1.86 -1.85 -1.83 -1.81 -1,77 -1,74 -1.69 St % * t St t t * * t » t t St t t t « St t $ t t --------------------- PATHS PAGE 5 10 RUN S EXI RETURN EARTH CURRENT IN TERMINAL GROUND SYSTEMS NAME t------t % T1 * T2 * t------- TERMINAL POSITION LEFT RIGHT MODULE AMPERES 6058.651 2390,222 ANGLE RADIANS ----------1 X -1.37 X -1.64 X X ----------i D-20 PATHS PAGE I 11 RUN I EXI 0, I__ 900, 1800, 2700, 3600, 4500. 3!?rn w s s c s a s :s: o » -s —io m c n 0,0000-1 I I I 1 1 I I I I 7,5000-1 I I I I I I I 1 I . CURRENT IN EACH SECTION LEFT SIDE OF FAULTED TONER D-21 5400. CURRENT (HOMLE) AHPERES 6300. 7200. 8100. 9000, PATHS PAGE RUN t EXI l 12 CURRENT (NODULE) AHPERES 1200. 1800. 2400, 3000. 3600, 4200, 4800. 5400. 6000, 0. 600. I______ I_ _ _ _ __ J_____ „I_ _ _ _ _ _ _ I_ _ _ _ _ _ _ I_ _ _ _ _ _ _ I......... .... I_______ I_ _ _ _ _ _ _ I_______ I 0.-I 4 I « I » I I 1 I » I I » I * 10,-I t I I t 1 4 i i i 4 i « 4 i i 4 i 4 20,-1 4 I 4 I 4 I 4 I 4 1 4 I 4 I 4 I 4 I 4 30.-I 4 I 4 I 4 I 4 I 4 4 I I 4 4 I I 4 4 I 40,-1 4 4 I T l S E C T 1 0 N N U H B E R I 70,-1 I I I I 1 I I I I 80,-1 I I I I I 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 . CURRENT IN EACH SECTION RIGHT SIDE OF FAULTED TOUER D-22 PATHS PA6E *. 13 RUN 1 EXI O, 41, I__ I__ O.OOOO-I CURRENT (HOWIE) AHPERES 82. 123. 164. 205. 246. 287, 328, 369. 410, .1______ I______ I_______ I_______I_______I_______ I______ I_______I I I T 0 y E R 1 I I I I I I 7.0000-1 I N U H B E R I I I I I I I I » CURRENT IN EACH TOUER LEFT SIDE OF FAULTED TWER D-23 PATHS PA6E I 14 RUN 1 EX1 CURRENT (HOWIE) AMPERES 0, 40, 80, 120. 160. 200, 240, 280. 320, 360, 400. I___ ___ I_ _ _ _ _ _ _ I_ _ _ _ _ _ _ I_ _ _ _ _ _ _ I______ I______ I_______ I______ I_______ I_ _ _ _ _ _ _ I 0,-1 4 I 6 I 4 I ♦ I » I 4 I 4 I 4 I 4 I 4 10,-1 4 I 4 I 4 I 4 I 4 I 4 I 4 I 4 I 4 I 4 20,-1 4 I 4 I 4 1 4 I 4 I 4 I 4 I 4 I 4 I 4 30,-1 4 1 4 I 4 I 4 4 I l T 0 U E R I I I I ?0.-I I I I I I I I N U M 6 E R I I 80,-1 I I I I . CURRENT IN EACH TOWER RIGHT SIDE OF FAULTED TOWER END OF PROGRAM PATHS mmmtmmmt D-Zk APPENDIX E LIST OF SURVEY RESPONDENTS 1 - Alabama Power Company 2 - Alcan Smelter & Chemicals Ltd. 3 - Allegheny Power Service Corporation 4 - Atlantic City Electric Company 3 - Arkansas Power & Light Company 6 - Arizona Public Service Company 7 - American Electric Power Service Corporation 8 - Bonneville Power Administration 9 - Boston Edison Company 10 - Baltimore Gas & Electric Company 11 - British Columbia Hydro Power Authority 12 - Calgary Power Ltd. 13 - Consumers Power Company 14 - Cincinnati Gas & Electric Company 15 - Columbus & Southern Ohio Electric Company 16 - Central Hudson Gas & Electric Corporation 17 - Commonwealth Edison Company 18 - Central Illinois Public Service Company 19 - Central Louisiana Electric Company Inc. 20 - Delmarva Power & Light Company 21 - Duke Power Company 22 - Dallas Power & Light Company 23 - Florida Power Corporation 24 - Gulf States Utilities Company 25 - Georgia Power Company 26 - Hawaian Electric Company Inc. 27 - Hydro-Quebec 28 - Houston Lighting & Power Company 29 - Illinois Power Company 30 - Iowa Public Service Company 31 - Indianopolis Power & Light Company 32 - Iowa Electric Light & Power Company 33 - Idaho Power Company 34 - Kansas Gas & Electric Company 35 - Kansas City Power & Light Company 36 - Kentucky Utilities Company 37 - Los Angeles Department of Water & Power 38 - Lake Superior District Power Company 39 - Long Island Lightning Company 40 - Missouri Utilities Company 41 - Metropolitain Edison Company 42 - Minnesota Power & Light Company 43 - Mississippi Power & Light Company 44 - Nevada Power Company E-1 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 - New Brunswick Electric Power Commission New Orleans Public Service Inc. Northwestern Public Service Company Nebraska Public Power District Northeast Utilities Service Company Niagara Mohawk Power Corporation New York State Electric & Gas Corporation Nova Scotia Power Corporation Ontario Hydro Ohio Edison Company Otter Tail Power Company Omaha Public Power District Orange & Rockland Utilities Incorporation Public Service Company of Oklahoma Pensylvania Power Company Puget Sound Power & Light Company Public Service Electric & Gas Company Pensylvania Power & Light Company Puerto Rico Water Ressources Authority Philadelphia Electric Company Portland General Electric Company Pacific Gas & Electric Company Pacific Power & Light Company Potomac Electric Power Company Rochester Gas & Electric Corporation Southwestern Public Service Company Southern California Edison Company Saskatchewan Power Corporation St. Joseph Light & Power Company San Diego Gas & Electric Company Salt River Project Agricultural Improvement & Power District Sierra Pacific Power Company Seattle City Light Department Tampa Electric Company Texas Electric Service Company The Dayton Power & Light Company Toledo Edison Company Tennessee Valley Authority Utah Power & Light Company Virginia Electric & Power Company Wisconsin Electric Power Corporation Wisconsin Public Service Corporation Wisconsin Power & Light Company Washington Water Power Company. E-2 APPENDIX F ROCHESTER GAS AND ELECTRIC CORPORATION (RGE) 115 KV TRANSMISSION LINE GROUNDING \* At ics. j TEST Figure F.1 TEST RESULTS RGE1s 115 kV Transmission Line Test Site F- 1 Kysorvi RESISTIVITY MEASUREMENTS TRAVERSE i ^ 01-1 TRAVERSE Traverse parallel to transmission line. Center of traverse about 60 meters (200 feet) north of structure 161. I Wenner method i Equipment used: Megger RESISTIVITY MEASUREMENTS TABLE f- I METER READING c Traverse perpendicular to transmission line. Center of traverse about 60 meters (200 feet) north of structure 161. 9 Wenner method i Equipment used: Megger MEASURED VALUES SCALE FACTOR s APPARENT RESISTANCE R = c. s (f!) 55.6 APPARENT RESISTIVITY P (Q-m) igST? 0.5 0.25 556 0.1 1.0 0.25 274 0.1 2.0 0.25 850 0.01 8.5 0.25 306 0.01 3-06 57-9 0.25 69 0.01 0.69 21.7 0.25 - 3.0 5.0 10.0 * 178.2 107.8 27.4 - NOTE - - ** METER PROBE PROBE READING SPACING DEPTH c a b (Meters (Meters) E = 4 + aVb1 b /F ; APPARENT RESISTANCE R - c.s APPARENT RESISTIVITY P (ft-m) (a) NOTE 1.0 0.25 266 0.1 173.0 * 2.0 0.25 981 0.01 9-81 124.4 * 0.25 595 0.01 5.95 112.6 0.25 367 0.01 3.67 115.5 0.25 196 0.01 1.96 123.2 3.0 5.0 10.0 16.6 * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. 2trbR P = -----------------------------—--------—------2 In (2 * + 2/F - /t - £ 1 + SCALE FACTOR s ** * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. where 01-2 9 MEASURED VALUES PROBE PROBE SPACING DEPTH a b (Meters) (Meters) rge TABLE F = 1 + a2/bJ ** Questionable readings due (very probably) to buried metal structures. 2*rbR p = ------------ -- ----------------------- ------------21n(-.-...—) + 2/F - /E - f where 1 . £ - k + az/bz b ; F - 1 + a2/b2 f-2 RESISTIVITY MEASUREMENTS TRAVERSE rge RESISTIVITY MEASUREMENTS TABLE TRAVERSE 03 i Traverse parallel to transmission line and adjacent to pole 162. Center of traverse about 6 meters. 1 Wenner Method i Equipment used : Megger o<, I Traverse parallel to transmission line and along path between poles 161 and 160. Center of traverse about halfway between poles. I Wenner method 8 Equipment used : Megger MEASURED VALUES MEASURED VALUES METER PROBE PROBE READING SPACING DEPTH c a b (Meters (Meters) TABLE SCALE FACTOR s APPARENT RESISTANCE R * c.s (0) APPARENT RESISTIVITY P METER PROBE DEPTH READING b c (Meters) NOTE PROBE SPACING a (Meters) * 0.5 0.1 635 1.0 0.1 106 (Q-m) APPARENT RESISTANCE R = c.s (n) APPARENT RESISTIVITY 0.1 63.5 204.0 0.1 10.6 67.0 SCALE FACTOR s P (Q-m) 1.0 0.25 236 0.1 23-6 153-5 2.0 0.1 117 0.1 11.7 147.2 3.0 0.1 606 0.01 6.06 114.3 2.0 0.1 481 0.01 4.81 60.5 5.0 0.1 236 0.01 2.36 74.2 ** 3.0 0.1 436 0.01 4.36 82.2 10.0 0.1 70 0.01 0.70 44.0 ** 4.0 0.1 392 0.01 3.92 98.5 30.0 0.1 - ** 5.0 0.1 335 0.01 3.35 105.3 145.8 - - - * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. 8.0 0.1 290 0.01 2.90 12.0 0.1 249 0.01 2.49 187.7 16.0 0.1 214 0.01 2.14 215.1 20.0 0.1 118 0.01 1.18 148.3 30.0 0.1 57 0.01 0.57 107.4 2irbR 27TbR p * E *= A + a 2/b* b /f ; f - 1 + a 2/b2 ** Unreliable readings due (very probably) to buried metal structures. ------------------------------------------------------------------------ 2 1n(2 * ■£) . 2/F - /I - £ 2ln(IjLj!) ♦ 2/F - /E - £ 1 + * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. p * --------- -------------------where NOTE where 1 ♦ /F E => It + aJ/b2 b ; F = 1 + a2/b2 POTENTIAL MEASUREMENTS PROFILE I 8 i I rge TABLE f-5 POTENTIAL MEASUREMENTS PROFILE 02-1 Structure 161, pole A. Low Frequency Test Return current electrode about 300 meters feet) north of transmission line right of way Overhead ground wires connected to transmission line grounds Equipment used: - Selective voltmeter + ammeter - Power Frequency generator. B l (985 I I CURRENT I (Amperes) 2.0 20.0 40.0 60.0 2.0 80.0 100.0 2.0 2.0 120.0 llto.o 2.0 2.0 160.0 180.0 2.0 2.0 200.0 220.0 2.0 2.0 2140.0 260.0 276.0 Return Current Probe 2.0 2.0 2.0 2.0 RGE 02"2 MEASURED VALUES VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /r2 -4- (J|2L2 <fi) 20.0 f-6 Structure 161, pole B, Low Frequency Test Return current electrode about 300 meters (985 feet) north of transmission line right of way Overhead ground wires connected to transmission line grounds Equipment used: - Selective voltmeter + ammeter - Power frequency generator MEASURED VALUES POTENTIAL PROBE LOCATION (Meters) TABLE 4.5 4.9 4.5 4.5 4.5 4.7 4.7 4.7 4.6 4.5 4.8 4.6 4.7 5.1 2.0 5.3 2.0 46.6 MEASURED POTENTIAL CURRENT (in % NOTE FREQUENCY of GPR ) (Hertz) 2.25 50 2.49 2.25 2.25 100 2.25 2.35 50 50 2.35 2.35 50 50 2.3 2.25 50 50 2.4 2.3 50 50 2.35 2.55 50 50 2.65 50 23.3 50 50 POTENTIAL PROBE LOCATION (Meters) 3.35 0. 1.0 2.5 5.0 10.0 20.0 60.0 150.0 210.0 240.0 CURRENT I (Ampe res) VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /r2 + w2l2 (n) 2.0 2.0 1.7 0.85 0.0 0.0 2.0 2.0 5.5 5-6 2.75 2.8 2.0 5.0 2.0 2.0 5.6 5.7 2.0 2.0 2.0 2.0 6.2 5.9 5.9 2.5 2.85 0! 3.1 2.95 2.95 6.2 3.1 . MEASURED POTENTIAL CURRENT (in % FREQUENCY of GPR ) (Hertz) 50 50 50 50 50 50 so 50 50 50 50 50 * Measurement made at pole A of structure 161 ** Measurement made at pole B of structure 161 (origin of profile) NOTE * ** POTENTIAL MEASUREMENTS PROFILE RGE 1 I I I POTENTIAL MEASUREMENTS TABLE F PROFILE °2-3 I I Structure 161 pole A, Frequency Sweep Test Return Current electrode about 300 meters feet) north of transmission line right of way Overhead ground wires connected to transmission line ground Equipment used: - Selective voltmeter + ammeter “ Power frequency generator (985 I I IhO.O CURRENT I (Amperes) (985 240.0 240.0 240.0 2.0 2.0 240.0 2.0 2.0 2.0 240.0 240.0 24o.o 240.0 240.0 240.0 --------- 241770----- MEASURED VALUES VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /r* + w2L2 (!5) 2.0 2.0 2.0 2.0 1.0 1.0 . s 0 47 RGE 02-4 Structure 161 pole B,Frequency Sweep Test Return current electrode about 300 meters feet) north of transmission line right of way Overhead ground wires connected to transmission line grounds Equipment used: - Selective voltmeter + ammeter - Fewer frequency generator MEASURED VALUES POTENTIAL PROBE LOCATION (Meters) TABLE 4.7 5.1 5.3 5.8 6.3 7.0 8.5 10.2 12.5 7.1 9.6 MEASURED POTENTIAL CURRENT (in % FREQUENCY of GPR ) (Hertz) 2.35 so 2.55 2.65 7R 100 2.9 3.15 3.5 4.25 5.1 6.25 7.1 9.6 12.0 NOTE 240.0 240.0 240.0 240.0 150 200 250 400 600 1000 1000 POTENTIAL PROBE L0CAT1 ON (Meters) * * 2000 4971 * Difference between the reading is probably caused by the sine wave distorsion of the power frequency generator at 2 amperes and frequencies above 1 KHz. CURRENT I (Amperes) VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /r2 + oj2L2 (12) 240.0 240.0 2.0 2.0 2.0 2.0 2.0 1.0 240.0 0.496 6.2 6.7 7.7 8.6 13-0 9.2 4.6 3-1 3.35 3.85 4.3 MEASURED POTENTIAL CURRENT (i n % NOTE FREQUENCY of GPR ) (Hertz) 50 6.5 9.2 100 200 400 1000 2000 9-274 5000 POTENTIAL MEASUREMENTS PROFILE i fi B B TABLE f-9 POTENTIAL MEASUREMENTS PROFILE RGE 02-4 Structure 159 pole B, Frequency Sweep Test Return current electrode about 300 meters ($85 feet) north of transmission line right of way Overhead ground wires connected to transmission line grounds Equipment used: - Selective voltmeter + ammeter - Power frequency generator 8 1 B I 2i(0.0 2*40.0 CURRENT I (Amperes) MEASURED VALU ES VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /r2 + u)2L2 (a) 240.0 2.0 2.0 2.0 240.0 240.0 2.0 2.0 240.0 2.0 02-5 Structure 160 pole B, Frequency Sweep Test Return current electrode about 300 meters (985 feet) north of transmission line right of way Overhead ground wires connected to transmission line grounds Equipment used: - Selective voltmeter + ammeter - Power frequency generator MEASURED VALUES POTENTIAL PROBE LOCATION (Meters) rge TABLE 4.6 5.15 6.45 8.60 13.0 20.6 2.3 2.575 3-225 4.3 6.5 10.3 MEASURED POTENTIAL CURRENT (in % NOTE FREQUENCY of GPR) (Hertz) POTENTIAL PROBE LOCATION (Meters) CURRENT I (Amperes) 50 100 200 240.0 240.0 240.0 400 1000 240.0 240.0 2.0 2.0 2.0 2.0 2000 240.0 VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /r2 + ai2L2 (17) MEASURED POTENTIAL CURRENT (i n % FREQUENCY of GPR) (Hertz) 2.0 3.35 3.95 5.25 7.70 13.2 1.675 1.975 2.625 3.85 6.6 50 100 200 400 1000 2.0 18.6 9-3 2000 NOTE POTENTIAL MEASUREMENTS PROFILE RGE g i I 1 POTENTIAL MEASUREMENTS TABLE p-n PROFILE °2-7 Structure 162 poles A and B, Frequency Sweep Test Return current electrode about 300 meters (985 feet) north of transmission line right of way Overhead ground wires connected to transmission line grounds Equipment used; - Selective voltmeter + ammeter - Power frequency generator I I I I I MEASURED VALUES VOLTAGE V (Volts) APPARENT IMPEDANCE z = A2 * u>2l2 (fi) A.9 MEASURED POTENTIAL CURRENT (in * FREQUENCY of GPR) (Hertz) NOTE 2.45 2.65 3.2 4.05 50 100 200 400 A A A A 210.0 2.0 2.0 2.0 210.0 2.0 5-3 6.4 8.1 210.0 210.0 2.0 2.0 17.7 6.35 8.85 1000 2000 A A 210.0 210.0 2.0 2.0 7.7 7.6 3-85 3.8 50 100 B B 210.0 2.0 14.3 7.15 1000 B 210.0 210.0 ■n CURRENT 1 (Amperes) 12.7 CO A = pole A B = pole B 02-8 Structure 163» Frequency Sweep Test Return current electrode about 300 meters (985 feet) north of transmission line right of way Overhead ground wires connected to transmission line grounds Equipment used; - Selective voltmeter + ammeter - Power frequency generator MEASURED VALUES POTENTIAL PROBE LOCATION (Meters) rge TABLE F-'2 POTENTIAL PROBE LOCATION (Meters) CURRENT 1 (Amperes) VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /r2 + iu2L2 (n) 210.0 210.0 2.0 2.0 2.3 2.8 1.15 1.4 210.0 2.0 4.05 2.025 210.0 210.0 2.0 2.0 6.25 11.0 210.0 210.0 2.0 2.0 16.3 25.6 3.125 5.5 8.15 12.8 MEASURED POTENTIAL CURRENT NOTE (in % FREQUENCY of GPR ) (Hertz) 50 100 200 400 1000 2000 5000 POTENTIAL MEASUREMENTS PROFILE I i 1 I RGE TABLE f->3 POTENTIAL MEASUREMENTS 02-9 PROFILE Structure 160 poles A and B, Frequency Sweep Test Return current electrode about 300 meters feet) north of transmission line right of way Downleads disconnected from pole ground rods Equipment used: - Selective voltmeter + ammeter - Power frequency generator i i (985 I i CURRENT 1 (Ampe res) 02-10 Structure 161 poles A and B, Low Frequency Test Return current electrode about 300 meters (985 feet) north of transmission line right of way Downleads disconnected from pole ground rods Equipment used : Megger MEASURED VALUES POTENTIAL PROBE LOCATION (Meters) rge TABLE MEASURED VALUES VOLTAGE V (Volts) APPARENT IMPEDANCE 2 = /r2 . gj2 L2 (a) ho MEASURED POTENTIAL CURRENT (i n % FREQUENCY of GPR ) (Hertz) 50 NOTE 200.0 20370 1.0 1.0 ')0.0 5071 50.1 1000 A A 200.0 200.0 1.0 1.0 33.7 33-7 33.7 33.7 50 100 B B 200.0 200.0 200.0 1.0 1.0 1.0 33.5 33.1 32.7 33.5 33.1 32.7 200 500 1000 B B B 200.0 1.0 32.6 32.6 2000 B 200.0 200.0 1.0 1 .0 20.3 20.3 19.8 SO 1000 AB AB 19.8 POTENTIAL PROBE LOCATION (Meters) 200.0 200.0 METER READING SCALE FACTOR 828 0.1 1.0 APPARENT IMPEDANCE 2 Z = /r 22 . g L POLE ID {a) 107 82.8 107.0 MEASURED CURRENT FREQUENCY NOTE (Hertz) A B i Notes: A = pole A ground rod ; b = pole B ground rod AB = pole A ground rod in parallel with pole B ground rod. ( I APPENDIX TENNESSEE VALLEY AUTHORITY (TVA) 500 KV STAGED FAULT AND TOWER GROUNDING TEST RESULTS ,Portlano ClARKSVIi Gallatu iGoodlelis ' Xil TEST SITE 500 kV kj0'®. TRANSHISS IOli I I MC kflid Hii • i'3 \ Figure G.1 TVA1s 500 kV Johnsonvi11e-Cumber1 and Transmission Line G-1 ( I RESISTIVITY MEASUREMENTS RESISTIVITY MEASUREMENTS TABLE G-i TRAVERSE TRAVERSE (Measured by TVA) 3 S I 8 i Traverse parallel to transmission line and directed toward Johnsonvi11e Wenner method with one current electrode kept in same location (at tower 130) Equipment used: Megger MEASURED VALUES £T> l MEASURED VALUES APPARENT RESISTIVITY P (ft-m) APPARENT RESISTANCE (n) NOTE PROBE SPACING a (feet) 1.5 3.0 ItO.O 235.3 * 1.5 3.0 3.0 31.0 2it8.4 * 5.0 3.0 17-5 196.5 * 3.0 5.0 10.0 3.0 7.5 150.7 * 10.0 15.0 3.0 4.9 143.9 * 15-0 20.0 3.0 4.2 162.9 * 20.0 30.0 3.0 2.6 150.2 30.0 ko.o 3.0 1.55 119.1 40.0 50.0 3.0 1.3 124.7 50.0 60.0 3.0 1.1 126.6 60.0 70.0 3.0 0.9 120.8 70.0 100.0 3.0 1.1 191.6 100.0 * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. PROBE DEPTH b APPARENT RES I STANCE (n) APPARENT RESISTIVITY P (ft-m) (feet) 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 264.7 * 41.0 328.6 * 30.0 337.0 * 13.8 277.3 * 8.0 235.0 * 4.9 190.1 * 3-3 190.7 2.49 191.4 2.1 201.5 2.3 2.35 264.7 315.4 1.1 210.8 For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. ZtrbR -----------------------—------------------------------------- 2 in where p = E = i, + a2/b2 b ; F " 1 + a2/b2 --------------------------------------------- ------------ —- 2 In(2 * l/t) + 2/F - /E - £ + 2/F - /E - £ 1 ♦ /F NOTE 45.0 2nbR p = 1-2 tva (Measured by TVA) i Traverse parallel to transmission line and directed toward Cumberland Wenner method with one current electrode kept in same location (at tower 130) Equipment used : Megger PROBE PROBE SPACING DEPTH a b (feet) (feet) TABLE where 1 + /F E 3 A + a2/b2 b ; F = 1 * a2/b! g-2 RESISTIVITY MEASUREMENTS TRAVERSE tva RESISTIVITY MEASUREMENTS TABLE 2-1 TRAVERSE (Measured by TVA) Traverse perpendicular to transmission line in south-east direction i Wenner method with one current electrode kept in same location (at tower 130) Equipment used : Megger S Traverse perpendicular to transmission line in north-west direction B Wenner method with one current electrode kept in same location (at tower 130) I Equipment used: Megger MEASURED VALUES PROBE PROBE SPACING DEPTH a b (feet) (feet) 1.5 3.0 5.0 10.0 15.0 20.0 30.0 40.0 50.0 60.0 70.0 100.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 MEASURED VALUES APPARENT RESISTIVITY P (ft-m) APPARENT RESISTANCE (0) NOTE PROBE SPACING a (feet) 49.0 36.2 19.8 288.2 290.1 * 224.4 a 6.0 120.6 * 10.0 3.1 91.1 * 15.0 2.7 104.7 * 20.0 1.77 102.3 30.0 1.38. 106.0 40.0 1.2 115.1 50.0 120.1 60.0 124.8 70.0 153.3 100.0 1.05 0.93 0.8 * * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. 1.5 3.0 5.0 PROBE DEPTH b (feet) 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 APPARENT RESISTIVITY P (fi-m) APPARENT RESISTANCE (fi) 45.0 39.0 31.0 * 312.5 * 348.2 * 16.1 323.6 * 10.0 293.8 * 6.9 3.2 2.56 267.7 * 2.09 200.5 184.9 196.7 1.85 1.52 212.9 204.0 210.8 1.1 * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. 2TbR ------ ----------------------------------------------------------- P 2 In (— ■*) + 2/F - /E - £ 2 1n(2-—-) * 2/F - /E - £ where 1 ♦ /F £ = 4 + a2/b2 b ; F = 1 + a2/b2 NOTE 264.7 27TbR p * 2-2 (Measured by TVA) i I ™ TABLE ^ where 1 + /F E = 4 ♦ a2/b2 b ; F RESISTIVITY MEASUREMENTS TRAVERSE TABLE g-5 RESISTIVITY MEASUREMENTS TRAVERSE TVA 03 (Measured by SES) TABLE tva oa (Measured by SES) B Traverse orthogonal to transmission line in north-west direction 9 Traverse parallel to transmission line and centered at tower 130 I Wenner method with one current electrode kept in same location (at tower 130) i Wenner method with center of traverse kept at center of tower 9 Equipment used: 9 Current frequency kept constant at 70 Hz 9 Equipment used: - Selective voltmeter + ammeter - Power frequency generator - Selective voltmeter + ammeter - Power frequency generator MEASURED VALUES PROBE SPACING a (feet) O I U1 PROBE DEPTH b (feet) MEASURED VALUES APPARENT CURRENT VOLTAGE RESISTANCE V (P -P ) I(C1-C2) (fi) (amperes) (volts) 1 2 APPARENT RESISTIVITY P (fi-m) NOTE PROBE SPACING a (feet) 733. 682. 954. 297. * 2.0 * * 5.0 8.0 * 10.0 3.80 0.100 0.100 254.0 234.0 327.0 50.5 38.0 304.5 * 16.6 3-0 2.03 0.100 20.3 228. * 23-3 10.0 3-0 0.617 0.100 6.17 12**. * 30.0 * 1.5 0.15 2.54 0.010 1.5 0.25 2.34 0.010 1.5 0.25 3.27 0.010 1.5 3.0 5-05 3.0 3.0 5.0 15.0 3.0 0.657 0.200 3.285 20.0 3.0 0.558 0.200 2.79 96.5 108.2 30.0 3.0 0.390 0.200 1.95 112.7 90.0 1(0.0 3-0 0.286 0.200 1.43 109.9 110.0 1.26 120.9 150.0 130.6 160.9 50.0 3.0 0.252 0.200 60.0 3.0 0.193 0.170 70.0 3.0 0.168 0.180 1.13 0.933 100.0 3.0 0.126 0.150 0.84 40.0 70.0 PROBE DEPTH b (feet) 0.33 0.33 0.33 0.33 3.0 3.0 3.0 3.0 3.0 3.0 3.0 3.0 VOLTAGE CURRENT APPARENT V (P -P ) I (C1-C2) RESISTANCE (volts) (amperes) (fi) 1 2 1.780 0.650 0.443 0.025 71.2 26.0 17.72 0.200 10.25 0.200 5.30 1.06 0.395 0.308 0.150 2.63 277.0 249.6 276.8 206.0 171.6 118.6 0.150 2.05 118.6 0.242 0.150 1.61 124.0 0.164 0.130 1.26 169.3 0.140 0.130 1.077 185.7 0.151 0.150 1.01 212.2 0.169 0.200 0.845 242.8 0.025 0.025 2.05 formula must be used to determine the apparent resistivity. 27TbR ------------------------------- —------------------- ---------- p = 2 in (i?--!-*!) + 2/F - /E - £ 1 W? E « <l ♦ a2/b2 b ; * * * * * A * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. 2TTbR whe re NOTE 125.2 * For short spacings and relatively long buried probes, the following p = APPARENT RESISTIVITY P (fi-m) F = 1 ♦ a2/b2 ------------------------------------------------------------- — 2 In ♦ 2/F - /E - £• b t + /F where E = A ♦ a2/b2 ; F » 1 + a2/b2 G-6 RESISTIVITY MEASUREMENTS TRAVERSE TABLE TVA 10 {Measured by SES) i Traverse parallel to transmission line and directed toward Johnsonvi1 le i Wenner method with one current electrode kept in same location (at tower 130) I Equipment used: - Selective voltmeter ♦ ammeter - Power frequency generator MEASURED VALUES PROBE PROBE SPACING DEPTH b (feet) (feet) 1.5 0.25 CURRENT VOLTAGE APPARENT V (pi-p2) I (C1-C2) RESISTANCE (amperes) («) (volts) APPARENT RESISTIVITY P (fi-m) NOTE 2.86 0.02 11(3.0 417.3 * 45.0 265-0 * 1.5 3.0 9.01 0.20 3.0 3.0 8.1(2 0.20 42.1 337.4 * 5.0 3.0 1(.59 0.15 30.6 343-7 * 10.0 3-0 2.17 0.15 14.46 290.7 A 15-0 3.0 1.272 0.15 8.48 249.1 * * 20.0 3.0 0.7i(it o.u 5.31 206.1 30.0 3.0 0.510 0.11| 3.64 210.5 1(0.0 3-0 0.1(20 0.15 2.80 215-2 50.0 3.0 0.396 0.15 2.64 253.3 60.0 3.0 0.520 0.20 2.60 299.2 70.0 3.0 0.539 0.20 2.695 361.7 100.0 3-0 0.269 0.20 1.345 257.7 * For short spacings and relatively long buried probes, the following formula must be used to determine the apparent resistivity. 2trbR p = ------------------------------------------------------------------- 2 In where + 2/f - /E - £ 1 ♦ /F £ * A + a2/b2 G“6 b ; F = 1 ♦ a2/b2 G-7 POTENTIAL MEASUREMENTS PROFILE tva TABLE g-8 POTENTIAL MEASUREMENTS 01 (continued) Return current electrode in first gulley (- 315') I Profile orthogonal to transmission line in north-west direction i Overhead ground wires connected to tower 130 and to adjacent towers on each side of tower 130 I Equivalent tower impedance - 1.382 Q MEASURED VALUES CURRENT I (Ampe res) MEASURED VALUES VOLTAGE V (Volts) APPARENT IMPEDANCE Z * /r2 + w2L2 (n) MEASURED POTENTIAL CURRENT (in % NOTE FREQUENCY of GPR) (Hertz) POTENTIAL PROBE LOCATION (feet) CURRENT VOLTAGE i V (Ampe res) (Volts) 70 70 70 215.0 215-0 215.0 35.5 35.1 70 70 215.0 215-0 1.516 1.50 1.50 0.910 36.1 31.2 70 70 215-0 1.50 3-65 0.963 1 .015 1.056 30.3 26.5 1.20 6.35 5-85 5.60 1.527 1 . 107 23.7 19.9 70 70 70 719.0 215-0 215.0 1.50 1.67 1.55 1.61 1.69 0.195 0.185 0.96 0.965 SO 55.0 1-527 1.527 1.76 Oo 5570 1.527 1.825“ 1.153 1.179 1.195 16.6 16.7 ~TT3------ 75.0 85.0 1.527 1.527 11.4 95.0 105.0 1.527 1-527 1.225 1.251 1.266 135.0 155.0 165.0 195.0 1.527 1.527 0.851 0.866 0.878 0.891 0.897 21.0 26.0 1.527 1.527 .... 1-39 1.65 25.0 30.0 35.0 5573 1.527 1.527 1.527 ’•527 1.527 19.5 20.0 1.527 1.87 0.950 70 70 1.303 1.310 70 70 1.316 1.316 6.8 6.8 70 70 2.01 1.316 6.8 70 2,01 1.316 1.297 6.8 6.1 70 70 1.250 - 20 1.250 - 40 65 1.527 1.277 1.286 ... 1,527 1.527 1-99 2.00 1.527 1.527 2.01 2.01 215-0 1.527 225.0 315.0 1.527 1.527 215.0 0-500 215.0 215-0 1.00 0.625 1.25 1.50 70 70 70-------- 9.5 8.5 7.6 7.1 5.7 5.2 1-91 1.93 1.95 1.96 115.0 70 1.98 1.92 1.280 - 70 70 70 215.0 215.0 APPARENT IMPEDANCE Z = /r2 + iii2L2 (n) 38.6 37.5 36.6 1.30. 1.32 1.36 1.36 1.37 15-0 16.5 18.0 g-9 PROFILE TVA01 1 POTENTIAL PROBE LOCATION (feet) TABLE 1.50 1.50 1.50 1-92 1.97 2.23 2.65 3-20 MEASURED POTENTIAL CURRENT (in % FREQUENCY of GPR ) (Hertz) 1.280 - 70 1.313 - 100 1.671 1.767 - 200 - 600 2.133 2.30 - 1000 2.90 3.90 6.66 6.82 - 2000 5000 10000 - 10000 20000 5.216 _ 800 NOTE TABLE POTENTIAL MEASUREMENTS PROFILE tva POTENTIAL MEASUREMENTS 2-10 02 PROFILE I 9 Return current electrode in first gulley (- 315') I Profile along transmission line toward Johnsonvilie B Overhead ground wires connected to tower 130 and to 10 adjacent towers on each side of tower 130 I Equivalent tower impedance =* 1.382 Si Return current electrode in first gulley (- 3151) Leg to leg profile I Overhead ground wires connected to tower 130 and to 10 adjacent towers on each side of tower 130. B Equivalent tower impedance - 1.382 MEASURED VALUES CURRENT VOLTAGE V I (Amperes) (Volts) APPARENT IMPEDANCE Z = /r2 + oj2 L2 to) MEASURED POTENTIAL CURRENT (in % frequency of GPR ) (Hertz) 30.0 1.527 1.52 1.00 27.6 35.0 1)5.0 1.527 1.527 1.61 1.76 1.05 1.15 2l*.0 55.0 1-527 1.88 1.23 11.0 60.0 65.0 75-0 1.527 1.527 1.527 1.91* 2.00 2.10 1.27 1.31 1.38 8.1 5.2 85.0 1.527 1.527 2.17 2.20 1.1*2 1.1*1* -2.7 95-0 115.0 135.0 1.527 1.527 2.1*0 2.65 1.57 1.71* - 155.0 165.0 1.527 1.527 3.00 3-00 1.96 1.96 -1*1.8 -1*1.8 70 70 70 70 70 70 70 70 70 70 70 70 70 70 70 70 70 70 70 70 70 195.0 215.0 1.527 1.527 A.62 5-90 3.03 3.86 -119.2 - 179.3 70 70 225.0 C2 1.527 -229.2 70 15.0 1-527 1.26 0.83 16.5 1.527 1-527 1.527 1.28 1.30 0.81* 0.85 0.87 1.33 25.0 1.527 1.527 1.527 1.527 18.0 19.5 20.0 21.0 21).0 1.527 G-11 n i MEASURED VALUES POTENTIAL PROBE LOCATION (feet) tva TABLE 1.33 1.35 1.1*1 1.1*3 6.95 11*7.9 39-9 39.2 38.5 37.0 37.0 36.3 33.1* 32.0 0.87 0.88 0.92 0,91* 1*. 55 96.85 16.8 \.k -h.2 13.6 25.9 6908. NOTE (feet) POTENTIAL PROBE TO LEFT LEG RIGHT LEG 1.0 1.5 3.0 CURRENT I (Amperes) VOLTAGE V (Volts) APPARENT IMPEDANCE Z = /R2 ■*■ U)2L2 to) MEASURED POTENTIAL CURRENT (in % FREQUENCY of GPR) (Hertz) 29 28.5 1.50 1.50 0.200 0.301* 0.13 0.20 90.6 85.5 27.0 0.480 0.32 76.8 26.0 1.50 Jt.O 6.0 25.0 21*.0 1.50 1.50 1.50 0.570 0.675 0.761 0.1*5 0.51 72.5 67.1* 63.1 1.50 1.50 1.50 0.970 1.10 1.18 0.65 0.73 0.79 53.0 1*7.2 1^78 1.50 1.23 1.26 . * 1.22 0.82 1*0.7 1.50 1.50 1.50 0.81* 0.83 0.81 39-2 1.50 1.50 1.16 1.10 1*1*. 3 1*7.2 1.50 1.50 1.50 1.50 0.980 0.77 0.73 0.65 0.51 0.1*0 0.320 0.265 0.31 0.21 77-6 81*.8 0.175 0.117 91.5 5.0 22.0 20.0 8.0 10.0 TSTo 12.0 16.0 15.0 11*.0 *.0 15.0 ii 16.0 17.0 19.0 13-0 11.0 20.5 9.5 22.5 21*.5 7.5 5.5 l*.5 25.5 26.5 27.5 28.0 3.5 2.5 2.0 1.50 1.50 28.5 1.5 1 Rh - 1 1 21 0.765 0.605 0.1*65 0.38 0.177 39.9 1*1.1* 53.0 63.1 71.1 87.2 70 70 70 70 70 70 70 70 75 70 70 70 70 70 70 70 70 70 70 70 70 70 NOTE POTENTIAL MEASUREMENTS PROFILE tva TABLE G-'2 POTENTIAL MEASUREMENTS 05 PROFILE TABLE ™ I Profile parallel to transmission line,toward Cumberland B Along direction parallel to transmission line, toward Johnsonville 1 Return current electrode in second gulley (* 600 feet) i Return current electrode in second gulley (- 600 feet) 8 Overhead ground wires disconnected from tower 130 B Overhead ground wires disconnected from tower 130 i Tower resistance - 7*1 ohms I Tower resistance - 7*1 ohms MEASURED VALU ES POTENTIAL PROBE LOCATION (feet) MEASURED VALUES yOLTAGE V (Volts) APPARENT IMPEDANCE 2 = /R^, oj2L2 (fl) MEASURED potential CURRENT (in % NOTE FREQUENCY of GPR) (Hertz) POTENTIAL PROBE LOCATION (feet) -15.0 0.7 2.8 4.0 43.66 0.0 15.0 0-7 0.7 2.82 2.74 4.03 3-91 43.26 44.87 18.0 0.7 0.7 2.76 2.80 3.94 4.0 44.47 43.66 19-5 20.0 OJI 0.7 2.85 2.86 4.07 4.08 42.65 42.45 70 70 70 70 70 70 70 21.0 2A.0 25.0 30.0 35.0 0,-7.............. 2.89 3.00 0.7 3.03 0.7 3.20 0.7 3.30 41.85 39.64 39.03 35.61 33.60 70 70 70 70 70 25.0 0.7 4.13 4.28 4.33 4.57 4.71 45.0 0.7 3.56 5.08 28.37 70 51.0 60.0 65.0 75.0 85.0 na ....0-7 0.7 0.7 0.7 5.83 3.96 4.01 4.15 4.22 5.47 5.66 22.94 20.32 70 70 70 70 60.0 65.0 75.0 95.0 0.7 4.26 6.08 14.28 7o 70 105-0 115.6 105.0 115.0 0-7 0.7 4.30 4.35 6.14 6.21 13.48 12.47 70 70 135-0 155.0 135.0 155.0 0.7 0.7 4.45 4.50 6.36 6.43 10.46 9.46 70 70 165.0 0.7 4.55 6.50 8.45 0.7 0.7 4.65 4.70 6.64 6.71 0.7 0.7 1.4 4.70 ___ - CURRENT I (Amperes) 195.0 215.0 225.0 31.5..0 315..0 ... G-13 5.73 5.93 6.03 19.31 16.50 15.09 ' 15.0 nrs 18.0 19.5 20.0 21.0 24.0 30.0 35.0 45.0 55.0 85.0 95-0 CURRENT I (Amperes) APPARENT IMPEDANCE VOLTAGE V (Volts) Z = /ft2 + oj2Lz (n) 2.40 2.42 4.0 43.66 43.19 0.6 0.6 2.45 2.49 2.50 4.08 4.15 4.17 42.49 41.55 0.6 2.55 4.25 40.14 0.6 0.6 0.6 2.64 2.66 4.40 4.43 4.70 38.03 37.56 33.80 4.93 30.52 25.82 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 0.6 4.03 os 2.96 3.16 3-30 5.27 5.50 3.36 3.42 5.60 5.70 3.55 5.92 3.60 3-66 6.0 0.6 3.70 —5T7E 0.6 6.10 6.17 15.49 14.08 70 70 13.14 70 70 ... -3.fe2..... 6.68 0.697 4.79 6.87 70 215.0 0.698 4.83 6.44 5.43 70 70 225.0 OgT- ----- Ol 6.92 315.0 0.7 6.71 5.43 70 4.75 6.78 4.43 70 9-39 6.71 5.53 70 * Return electrode located in first Gulley (- 315 feet) 70 70 70 7.357 4.66 195-0 5.15 21.13 19.72 16.67 05B 165.0 .29 70 22.53 8.45 6.57 5.97 3.21 2.54 2.00 6.50 6.63 70 70 70 70 70 70 70 70 70 70 rr774 3.90 " liTTTi 07 0.6 0.698 * MEASURED POTENTIAL CURRENT (in 1 FREQUENCY of GPR) (Hertz) 70 70 70 70 70 70 70 NOTE POTENTIAL MEASUREMENTS TABLE G-1'* POTENTIAL MEASUREMENTS PROFILE PROFILE TVA°7 I Profile perpendicular to transmission line, along north-west direction I Return current electrode in second gulley (- 600 feet) I Overhead ground wires disconnected from tower 130 I Tower resistance - 7*1 ohms. I Profile across tower and between legs Return current in second gulley (- 600 feet) l Overhead ground wires disconnected from tower 130 B Tower resistance ~ 7*1 ohms MEASURED VALU ES (Amperes) VOLTAGE V (Volts) CURRENT l TVA 08 i MEASURED VALUES POTENTIAL PROBE LOCATION (feet) APPARENT IMPEDANCE Z = /r2 + u>2 L2 (£2) MEASURED POTENTIAL CURRENT (in % FREQUENCY of GPR ) (Hertz) (feet) NOTE POTENTIAL PROBE TO LEFT RIGHT LEG LEG CURRENT 1 (Amperes) VOLTAGE V (Volts) APPARENT IMPEDANCE 2 2 -10 0.0 0.6 2.40 4.0 43-66 70 2.0 28,0 0.6 0.6 0.6 2.44 2.48 4.07 4.13 42.72 41.78 70 70 3-0 27.0 - 18.0 0.6 2.52 4.20 40.84 70 4.0 5.0 26.0 25.0 0.6 0.6 1.356 1.590 2.26 2.65 iq.q 20.0 0.6 0.6 2.56 2.58 4.27 39-90 39.44 70 70 24.0 23-0 0.6 0.6 1.800 1.940 3.0 21.0 0.6 2.62 4.37 38.50 70 n n.6 70 70 2.09 2.18 10.0 20.0 0.6 —J7S1 STTH 36.62 35.45 31.92 23717 0.6 0.6 0.6 0.6 0.6 4.50 4.58 4.83 22.0 21.0 25.0 30.0 35.0 2.70 2.75 2.90 6.0 7.0 8.0 9-0 70 70 70 70 45.0 0.6 3.16 5.27 25752 55.0 60.0 65.0 0.6 0.6 0.6 3.30 3.39 3.41 5.5 5.65 5.68 22.5 20.4 75.0 85.0 0.6 0.6 3-50 3.60 5-83 6.0 05.0 105.0 0.6 6.02 0.6 3.61 3.66 115.0 0.6 3.72 6.1 6.2 135.0 155.0 165.0 195-0 0.6 0.6 0.6 0.6 3.80 3.81 3-85 3.90 6.33 6.35 6.42 6.5 215.0 0.6 3.93 6.55 225.0 315.0 0.6 0.6 3.95 ... 4.06 6.58 6.77 19.95 17.84 15.49 70 yu 19.0 12.0 18.0 13.0 17.0 11.0! 1.44 ** 3.23 3-48 3.63 2.25 2.32 3.75 0.6 0.6 2.36 3.93 2.38 2.40 3.97 4.0 3.97 3.87 44.6o 44.13 43.66 70 70 70 70 70 70 70 70 70 70 70 70 70 43.66 44.13 70 70 70 70 70 70 n-i - 68.17 62.67 57.75 54.46 50.94 48.82 57rTg~ 45-54 14.0 16.0 70 16.0 17.0 14.0 13.0 0.6 0.6 2.40 2.38 12.0 11.0 10.0 9.0 0.6 0.6 2.35 2.30 3.92 0.6 0.6 2.25 2.19 3.75 3.65 44.83 46.01 47.18 48.59 0.6 0.6 2.10 3.5 3-3 3-07 50.70 53-52 56.8l 75 18.0 70 70 12.67 8755 70 70 70 70 7S 19-0 20.0 21.0 7.75 70 25.0 26.0 7.28 4.69 70 70 10.56 9.62 - MEASURED POTENTIAL CURRENT NOTE (in % FREQUENCY of GPR ) (Hertz) 0.6 0.6 15.26 14.08 10.80 0.865 22 Z = /r + U) L (£ ) 15.0 16.5 ■>U TABLE g-15 22.0 23.0 24.0 8.0 7.0 6.0 5.0 0.6 1.98 1.84 4.0 3.83 . 70 70 70 0.6 1.67 1.45 2.78 02 60.80 65.96 70 70 27.0 --3,0. 2.0 0.6 0.6 1.26 1 .02 2.10 1.70 70.42 76.06 70 70 29.0 0.6 0.675 1.125 84.15 70 4.0 28.0 Note 1.0 0.6 Presence of roots at surface of soil * POTENTIAL MEASUREMENTS PROFILE g TABLE g-i 6 TVA 09 Profile perpendicular to transmission line along south-east direction i Return current electrode in second gulley I Overhead ground wires disconnected from tower 130 600 feet) I Tower resistance - 7*1 ohms HEASURED VALUES POTENTIAL PROSE LOCATION CURRENT 1 VOLTAGE V (Amperes) (Volts) (feet) APPARENT POTENTIAL IMPEDANCE Z = A2 + a)2 L2 to) of GPR ) MEASURED CURRENT NOTE FREQUENCY (Hertz) 0.600 0.600 2.22 2.23 3.7 3.72 67.88 67.65 19-5 0.600 0.600 2.25 2.27 3-75 3.78 67.18 66.71 70 70 20.0 21.0 0.600 0.600 2.28 2.31 3.8 3-85 66.68 65.77 70 70 .0 25-0 30.0 0.600 0.600 0.600 2.1(0 2.1(5 2.66 6. 6.08 —6.63 63.66 62.69 373E 70 -------70-------- 15.0 16.5 18.0 2k 70 70 70 35.0 0.600 2.81 6.68 103 70 1(5.0 55.0 0.600 o.6oo 3.05 3.20 5.08 5-33 28.60 70 70 60.0 0.600 65.0 75-0 0.600 0.600 85.0 3.30 5-5 22.53 70 3.1(0 3.50 5-67 5-83 20.19 17.86 70 70 3.63 6.05 6.17 6.32 6.37 6.50 6.63 16.79 13.16 11.03 10.33 8.65 6.57 70 70 70 70 70 70 6.67 6.7 6.72 6.10 5-63 5.60 70 70 70 6.93 70 6.22 “TTS7 8765 -TOOT 95.0 105.0 115.0 135.0 155-0 0.600 0.600 0.600 0.600 0.600 3.70 3-79 165.0 195.0 215-0 0.600 0.600 k.00 0.600 6.02 6.03 225.0 0.600 6.05 6.75 315.0 165.0 iSSTo 0.600 0.600 0.600 6.08 6.01 J79S 6.8 165.0 165.0 0.600 0.600 3-80 3-70 6.33 6.17 Note * 3.82 3.90 3.98 6.6b 5750 Potential probe driven partially in a root G-11 10.80 fJTT? 70 1989 5000 * i i GROUND RESISTANCE MEASUREMENTS i Measurements made by TVA on October (during the staged fault test period) at 10 towers adjacent to tower 130 (5 on each side) I Tower number increases from JohnsonvHle to Cumberland I TABLE GROUND RESISTANCE MEASUREMENTS Equipment used: Megger TOWER NUMBER 126 (Q.) 4.0 53--------- 127 7.5 5-5 130 131 132 6.0 10.0 11.0 133 1311 10.5 15.5 135 q.o 128 129 TOWER NUMBER GROUND RESISTANCE (G) TOWER NUMBER GROUND RESISTANCE (Q) I Measurements made by TVA from May to July 1970 on the Johnsonvi11e-Cumberland 500 kV transmission line section ■ Tower number increases from Johnsonvi1le to Cumberland B Equipment used: Megger TOWER NUMBER GROUND RESISTANCE (G) 6 9 TOWER NUMBER GROUND RESISTANCE (Gl 10 h 5 20 29 6 22.5 12 36 37 9 15 9 67 68 17 69 16 9 13 12 70 71 11 6 72 73 21 2 3 7 8 9 10 11 12 13 14 15 16 17 4 20 14 5 14 16 40 16 39 40 41 42 43 44 45 66 22 10 20 18 2475 4 77 78 8 79 2 50 3 80 5 3 7 20 81 10 82 3 20 15 83 84 85 14 10 2 16 86 87 12 8 51 22 40 52 53 54 21) 63 64 65 15 21 13 7 23 9 9 3 6 4 48 49 46 47 19 20 23 6! 62 7 15 5 20 11 21 6 8 5 18 . TOWER NUMBER 31 32 33 34 35 1 2 i.75 GROUND RESISTANCE (.01 6 10 5 CO 125 GROUND RESISTANCE TABLE G-18 10 5 20 74 75 76 55 56 27 6 19 3 57 8 28 4 7 58 59 16 10 88 29 89 4 3 3.Q 5 60 8 90 is 25 26 5 TABLE GROUND RESISTANCE MEASUREMENTS (continued) I Measurements made by TVA from May to Juiy 1970 on the Johnsonville-Cumberland 500 kV transmission line section I Twer number increases from Johnsonvi lie to Cumberland i Equipment used: Megger TOWER NUMBER GROUND RESISTANCE (Q) GROUND RESISTANCE (ft) 5 15 151 152 GROUND RESISTANCE (ft) 3 1 153 154 3 5 6 155 156 127 9 157 7 3 9 128 7 8 6 158 159 160 12 161 162 11 21 14 163 16 164 165 166 15 23 1 TOWER NUMBER 91 92 6 5 121 122 93 95 96 22 20 123 124 125 10 91* 10 4 126 97 13 98 99 21 3 100 101 n 13 102 103 16 17 104 105 10 4 129 130 131 132 13 2k 133 134 12 20 8 135 7 106 4 136 6 107 108 7 5 137 138 9 12 109 3 12 110 111 3 19 139 140 112 113 m 20 6 12 141 142 143 3 15 7 115 12 116 117 118 6 18 119 5 11 120 6 13 144 8 145 146 147 8 148 149 13 11 L£2 6 7 1 G-14 TOWER NUMBER 4 4 g-,9