Sunteți pe pagina 1din 806

Effective December 6, 2006, this report has been made publicly available in accordance with Section 734.

3(b)(3) and published in accordance with Section 734.7 of the U.S. Export Administration Regulations. As a result of this publication, this report is subject to only copyright protection and does not require any license agreement from EPRI. This notice supersedes the export control restrictions and any proprietary licensed material notices embedded in the document prior to publication.

Power System and Railroad Electromagnetic Compatibility Handbook


Revised First Edition

Technical Report

Power System and Railroad Electromagnetic Compatibility Handbook


Revised First Edition
1012652

Final Report, November 2006

Cosponsors Association of American Railroads (AAR) 50 F Street NW Washington DC 20001 Project Manager M. Congdon American Railway Engineering and Maintenance-of-Way Association (AREMA) 8201 Corporate Drive, Suite 1125 Landover, MD 20785 Project Manager I. Alperovich

EPRI Project Manager B. Cramer

ELECTRIC POWER RESEARCH INSTITUTE 3420 Hillview Avenue, Palo Alto, California 94304-1338 PO Box 10412, Palo Alto, California 94303-0813 USA 800.313.3774 650.855.2121 askepri@epri.com www.epri.com

DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES


THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN ACCOUNT OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH INSTITUTE, INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE ORGANIZATION(S) BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM: (A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I) WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S CIRCUMSTANCE; OR (B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER (INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT. ORGANIZATION(S) THAT PREPARED THIS DOCUMENT Timerider Technologies, Inc. CorrComp R.G. Olsen Consulting ARC Technical Resources, Inc. James Stewart Consulting Union Switch and Signal

NOTE
For further information about EPRI, call the EPRI Customer Assistance Center at 800.313.3774 or e-mail askepri@epri.com. Electric Power Research Institute and EPRI are registered service marks of the Electric Power Research Institute, Inc. Copyright 2006 Electric Power Research Institute, Inc. All rights reserved.

CITATIONS
This report was prepared by: Electric Power Research Institute (EPRI) 3420 Hillview Avenue Palo Alto, CA 94304 Principal Investigator B. Cramer Timerider Technologies, Inc. 172 Summit Road Bishop, CA 93514 TimeriderTech@earthlink.net Principal Investigator M. House The authors of each chapter of this book are listed with the chapters. This report describes research sponsored by EPRI, Oncor Energy Delivery Services, The National Grid Transco Company, Association of American Railroads (AAR), and American Railway Engineering and Maintenance-of-Way Association (AREMA). The report is a corporate document that should be cited in the literature in the following manner: Power System and Railroad Electromagnetic Compatibility Handbook: Revised First Edition. EPRI, Palo Alto, CA, Oncor Energy Delivery Services, Dallas, TX, The National Grid Transco Company, Warwick, UK, Association of American Railroads (AAR), NW, Washington DC and American Railway Engineering and Maintenance-of-Way Association (AREMA), Landover, MD: 2006. 1012652.

iii

PRODUCT DESCRIPTION

This book is a source of technology and information for preventing and mitigating ac electrical interference problems on railroads. All aspects of electromagnetic compatibility (EMC) where railroad systems are the receptors are examined. This includes well-understood areas such as magnetic induction from transmission lines as well as less understood areas such as conducted interference from distribution systems and effects of harmonics. Chapters examine all known effects of ac interference, including personnel safety, operation of railroad equipment and systems, and damage to railroad equipment. Direction is provided for studying the effects of proposed new and upgraded installations, as well as root cause analysis of problems on existing installations. The handbook is structured for use either as a reference, or for in-depth study. To date, information on these issues has been incomplete and scattered. The need for a single reference document exists for both the railroad and power industries. This handbook would not have been possible without a strategic alliance with the Association of American Railroads (AAR) and the American Railway Engineering and Maintenance-of-Way Association (AREMA). It is the hope of all those involved in electrical issues of joint corridors that this work will help facilitate safe, problem-free operation of railroads and power systems. Results & Findings For the first time, one resource provides the tools necessary to understand and rectify ac interference problems on railroads. The tools provided can help personnel work backwards from effects to causes to find the best possible solutions. Savings can be significant. As several case studies demonstrate, problems that took weeks and months to track down can now be located in days or hours. Challenges & Objectives This book is specifically written to be easily understood and useful for a variety of people responsible for the reliability of railroad and power systems. The intended audience includes railroad personnel, power company personnel, EMC specialists, and equipment suppliers. These groups include electrical engineers, mechanical engineers, technicians, signal maintainers, regulators, and managers. By preventing ac interference problems through advance planning, by tracking down existing problems more quickly, and by applying mitigation that is both effective and economical, substantial savings can result. Through continued study and handbook updates, even greater economic benefits can be expected. The continuing alliance between AAR, AREMA, and EPRI will help promote safe and reliable movement both for electric power and for passengers and freight, providing a high-degree of safety for the general public.

Applications, Values & Use Periodic updates to this book will include developments in the railroad and power industries. Changes in signal systems and power systems have the potential to change the way these systems interact. By following the evolution of technology in future editions, this tool will continue to be effective. EPRI Perspective EPRI is in a unique position to accomplish the goals described in this book. Our past joint projects with AAR, and our network of experts in power delivery and electromagnetic compatibility have provided many of the tools included in this handbook. With continued support from AAR and AREMA, an extensive network of railroad personnel, railroad equipment suppliers, and railroad consultants has been added to the mix. The result is a unique opportunity to produce a truly useful tool. Approach The project teams first goal was to create a tool for solving any ac interference problem a railroad might experience and to do it in a way that any technician or engineer could apply. Another goal was to provide enough information to solve problems quickly while allowing for in-depth study. Keywords joint corridors ac interference induction inductive coordination electromagnetic compatibility (EMC) electromagnetic interference (EMI) railroad signaling communication and signaling

vi

ABSTRACT
This book is intended as a source of technology and data for the prevention and mitigation of ac electrical interference problems on railroads. Much of the information was developed through joint projects between the Electric Power Research Institute (EPRI) and the Association of American Railroads (AAR) over the past thirty years, and by technical committees of the AAR and the American Railway Engineering and Maintenance-of-Way Association (AREMA). All aspects of electromagnetic compatibility where railroad systems are the receptors are examined. This includes well-understood areas such as magnetic and electric induction from transmission lines, as well as less well-understood areas such as conducted interference from distribution systems and the effects of harmonics. Chapters examine all known effects of ac interference, including safety of personnel, operation of railroad equipment and systems, and damage to railroad equipment. Direction is provided for the study of the effects of proposed new and upgraded installations, as well as root cause analysis of problems in existing installations. The handbook is structured to be used as both a quick reference book and for in-depth study.

vii

AUTHORS

Project Manager Brian Cramer

Editorial Committee Irina Alperovich Ron Capan Mary Congdon Marvin Frazier Eddy Harrel Mike House Paul Jones Robert Olsen Mike Silva Robert Stevenson

Chapter 1:

Introduction and Background Brian Cramer and Mike House Fundamentals of Electromagnetic Compatibility (EMC) Jerry Ramie and James Stewart Overview of Power Systems James Stewart and Brian Cramer Overview of Railroad Signal Circuits Mike House and Ben Feely Abnormal Operation of Railroad Equipment Mike House and Ben Feely Damage to Railroad Equipment Mike House Personnel Safety Considerations Brian Cramer and Marvin Frazier Field Measurements Mike House and Brian Cramer Investigation Mike House and Brian Cramer ix

Chapter 2:

Chapter 3:

Chapter 4:

Chapter 5:

Chapter 6:

Chapter 7:

Chapter 8:

Chapter 9:

Chapter 10: Diagnostic Flow Chart Brian Cramer and Mike House Chapter 11: Graphic Evaluation of Proposed Changes Based on Simulations Marvin Frazier Chapter 12: Computer Modeling of Joint Corridors Marvin Frazier and Brian Cramer Chapter 13: Mitigation Options Brian Cramer, Marvin Frazier, and Mike House Chapter 14: Case Studies various contributors Chapter 15: Planning Guidelines for New/Upgraded Railroad Systems Brian Cramer Chapter 16: Planning Guidelines for New/Upgraded Power Systems Brian Cramer Chapter 17: Conclusions Brian Cramer Chapter 18: Glossary Mike Silva and Mike House Chapter 19: Errata Sheet

ACKNOWLEDGMENTS
In 1872 a patent was issued for electrical signaling apparatus for railroads. In 1893 commercial ac power transmission began in New York City. For most of the ensuing 110 years issues involving ac interference with railroad systems have been studied, debated, and mitigated with varying degrees of success. Expertise has been developed and forgotten several times during this period. And each new technology introduced has added another twist to the process of developing and maintaining compatible electrical systems. This Handbook is the result of the efforts of a great many people over a period of more than a century. At the core of this effort are a handful of people who have struggled to bring two industries that use electricity in very different ways to a common understanding. These people share a passionate belief that by understanding all of the technologies involved, the issues can effectively be addressed. You will see these peoples names listed as authors and editors. We thank the EPRI contractors who contributed their unique skill and technological ability. We also deeply appreciate the help of the EPRI utility participants who provided financial support. We owe special thanks to the Association of American Railroads (AAR) and the American Railroad Engineering and Maintenance-of-Way Association (AREMA) for their participation and support. Much of the work in this field for the last dozen years has been done by these two organizations and by their technical committees. These industry committees, and the volunteers who participate on them, have made immeasurable contributions to this field. Much of the information collected in this book is the product of these organizations. We would like to thank the members of the Project Advisory Committee (PAC). The PAC was tasked with keeping the Handbook on track. If the Handbook is the useful tool we envisioned, and if it is really understandable in its presentation, then the PAC deserves much of the credit. The PAC members are: Irina Alperovich, Norfolk Southern Railroad (AREMA) Mary Congdon, Canadian Pacific Railroad (AAR) Eddy Harrel, Oncor Energy Delivery Services Paul Jones, Earthing Risk Management (National Grid Company) The following individuals contributed photographs to the Handbook: Peter Brackett, Canadian Pacific Railroad Ron Capan, Union Switch & Signal Ben Feely, Union Switch & Signal Michael Herz, Pacific Gas & Electric Company xi

Mike House, Timerider Technologies Inc. Mike Silva, Enertech Dan Warren, Southern Company Dave Wright, Safetran Systems The following individuals provided reviews and editorial support: Irina Alperovich, Norfolk Southern Railroad Forrest Ballinger, GE Transportation Warwick Beech, Erico Richard Bowden, Burlington Northern Santa Fe Railroad Peter Brackett, Canadian Pacific Railroad Ron Capan, Union Switch & Signal Mike Choat, CSX Transportation Mary Congdon, Canadian Pacific Railroad John Danyluk, Comet Communications & Signal Ben Feely, Union Switch & Signal Marv Frazier, CorrComp Inc. Dave Gove, Transtector Eddy Harrel, Oncor Energy Delivery Services Mike House, Timerider Technologies Inc. Paul Jones, Earthing Risk Management Rod Leard, Saft America Bob MacMillan, CSX Transportation Fred Meeks, Burlington Northern Santa Fe Railroad Robert Olsen, Washington State University Jerry Ramie, ARC Technical Resources Mike Silva, Enertech Robert Stevenson, Consultant James Stewart, Consultant

Brian S Cramer EPRI Palo Alto, California

xii

HOW TO USE THIS HANDBOOK


Electromagnetic Interference (EMI) and Electromagnetic Compatibility (EMC) are complex fields of study. EMC work is often further complicated when the source of the interference and the receptor of the interference are very different technologies. Railroads and power systems are an example of this. While both industries have thousands of knowledgeable people, including a great many electrical engineers, few of them understand both industries. This handbook is specifically organized to fill this void. To achieve this goal, most chapters have been organized into three sections. They are: Introduction Quick-Start Version The quick-start version of each chapter is designed to allow the user to work effectively in the shortest time. It is possible to read the quick-start version of each relevant chapter and to begin to address a problem. Of course, the effectiveness of this method will depend on the users prior knowledge and experience. Detailed Version For those with a thirst for greater knowledge, and the time to satisfy it, read the detailed version of each chapter. In the end, this will benefit both you and your organization. It is highly recommended that every user read the detailed section of Chapter 2: Fundamentals of Electromagnetic Compatibility. A basic understanding of this field of study is necessary. The next thing to read is the chapter about the other industrys system. For example, if you are a power engineer, read Chapter 4: Overview of Railroad Signal Circuits preferably the detailed version. If you are a railroader, then we recommend you read Chapter 3: Overview of Power Systems. To effectively address interference issues, you need an understanding of EMC fundamentals and the basics of all systems involved. Throughout this handbook the authors and editors have tried to make each chapter and section stand-alone. You will see terms and acronyms defined repeatedly. This is done so the user can use the handbook as a reference with a minimum of flipping back and forth to fill gaps. Never the less, an extensive glossary is provided with both power and railroad terms.

xiii

CONTENTS

1 INTRODUCTION AND BACKGROUND ................................................................................1-1 Background ...........................................................................................................................1-2 What is AC Power Interference? ...........................................................................................1-4 Prior Programs ......................................................................................................................1-5 Terminology...........................................................................................................................1-6 Organization ..........................................................................................................................1-7 References ............................................................................................................................1-8 2 FUNDAMENTALS OF ELECTROMAGNETIC COMPATIBILITY (EMC) ..............................2-1 Introduction ...........................................................................................................................2-2 Quick-Start Version ...............................................................................................................2-2 Introduction to EMC..........................................................................................................2-2 Detailed Version ....................................................................................................................2-4 Introduction to EMC..........................................................................................................2-4 Interference Paths ............................................................................................................2-6 Differential and Common Mode Signal Types ................................................................2-15 Differential Mode Signal.............................................................................................2-16 Common Mode Signal ...............................................................................................2-17 Interference Types..........................................................................................................2-18 Common Mode Interference ......................................................................................2-18 Differential Mode Interference....................................................................................2-19 Receptor Circuit Unbalance .......................................................................................2-21 European and North American Models ..........................................................................2-23 European Model ........................................................................................................2-23 North American Model ...............................................................................................2-24 Personnel Safety ............................................................................................................2-25 Emissions & Immunity Testing .......................................................................................2-25 Types of Emissions Tests...............................................................................................2-26

xv

Types of Immunity Tests ................................................................................................2-26 Immunity Performance ...................................................................................................2-27 Regulatory Structure.......................................................................................................2-27 Basic EMC Standards ....................................................................................................2-28 EMC Standards in other Industries.................................................................................2-28 Legal Environments........................................................................................................2-31 Social Environments.......................................................................................................2-31 Addressing the gaps.......................................................................................................2-32 Summary ........................................................................................................................2-33 Railways Need To:.....................................................................................................2-33 Power Providers Need To:.........................................................................................2-33 Railway Signal Equipment Manufacturers need to: ...................................................2-34 References ..........................................................................................................................2-34 3 OVERVIEW OF POWER SYSTEMS......................................................................................3-1 Introduction ...........................................................................................................................3-2 Quick-Start Version ...............................................................................................................3-2 Introduction to the Power System.....................................................................................3-2 Detailed Version ....................................................................................................................3-4 Introduction to the Power System.....................................................................................3-4 Overhead Transmission Lines..........................................................................................3-8 Underground Transmission Lines...................................................................................3-12 Overhead Distribution Lines ...........................................................................................3-13 Primary Feeder Grounding ........................................................................................3-15 4-Wire Multi-Grounded Feeders............................................................................3-16 4-Wire Single-Point Grounded Feeders ................................................................3-17 3-Wire Ungrounded Feeders.................................................................................3-17 3-Wire Grounded (Unigrounded) Feeders ............................................................3-19 5-Wire Feeders .....................................................................................................3-20 Other Systems ......................................................................................................3-20 Secondary Feeder Grounding....................................................................................3-21 Other Considerations Related to Ground Currents....................................................3-22 Impact on EMC ..........................................................................................................3-23 Substations Near Railroad Tracks .............................................................................3-23 Underground Distribution Lines ......................................................................................3-24

xvi

Secondary Distribution ...................................................................................................3-25 4 OVERVIEW OF RAILROAD SIGNAL CIRCUITS ..................................................................4-1 Introduction ...........................................................................................................................4-2 Quick-Start Version ...............................................................................................................4-2 Detailed Version ....................................................................................................................4-3 Railroad Signaling Circuits Introduction.........................................................................4-3 Railroad Track Part of the Track Circuit ........................................................................4-4 Railroad Rails and Ties Mechanical and Electrical Properties ......................................4-5 Railroad Tracks............................................................................................................4-6 Ballast.............................................................................................................................4-13 Contaminants in Ballast..................................................................................................4-15 Railroad Track as an Electrical Transmission Line.........................................................4-16 Bondwires.......................................................................................................................4-17 Rails................................................................................................................................4-20 The Shunting Action of a Trains Axles...........................................................................4-20 Insulated Joints...............................................................................................................4-25 Insulated Joint Failure Modes.........................................................................................4-29 Track Circuits..................................................................................................................4-31 What Is (and What Isnt) a Track Circuit?.......................................................................4-31 Robinsons Patent ..........................................................................................................4-32 Railroad History ..............................................................................................................4-37 General Types of Signal Equipment...............................................................................4-38 What is a Signal System?...............................................................................................4-39 Signaled vs. Non-Signaled Trackage .............................................................................4-40 Signaled Trackage and its Signal Systems................................................................4-41 Signal Systems: Wayside Signals and Cab Signals.......................................................4-41 What is a Block?.............................................................................................................4-43 Signaling Basics .............................................................................................................4-44 What is an Interlocking? .................................................................................................4-45 Mechanical Interlocking .............................................................................................4-47 Power Interlocking .....................................................................................................4-47 Relay-Based Interlocking ...........................................................................................4-48 Solid-State Interlocking ..............................................................................................4-49 Types of Signaling Systems ...........................................................................................4-50

xvii

Wayside Signal Aspects and Indications........................................................................4-50 Colorlight Signals .......................................................................................................4-52 Crossing Flashers ......................................................................................................4-53 LED Colorlight Signals ...............................................................................................4-54 Searchlight Signals ....................................................................................................4-55 Fiber-Optic Searchlight Signals .................................................................................4-56 Position Light Signals ................................................................................................4-56 Semaphore Signals ...................................................................................................4-56 Signal Lighting Circuits ..............................................................................................4-57 Signal Lamps .............................................................................................................4-57 Cab Signal Systems .......................................................................................................4-58 Types of Signal Control Systems ...................................................................................4-58 Automatic Block Signals (ABS)..................................................................................4-58 Absolute-Permissive Block Signaling (APBS)............................................................4-60 Centralized Traffic Control (CTC) ..............................................................................4-61 Train Control Systems ...............................................................................................4-63 Mechanical Trip Stop .................................................................................................4-64 Intermittent Inductive Trip Stop ..................................................................................4-65 Indusi .........................................................................................................................4-66 Magnetic Trip Stop.....................................................................................................4-66 Communications-Based Train Control (CBTC) ..............................................................4-67 Track Circuits in General ................................................................................................4-68 Batteries and Electrification ............................................................................................4-69 Primary Batteries ............................................................................................................4-70 Rechargeable Batteries ..................................................................................................4-71 Transmitters....................................................................................................................4-72 Receivers Correct Reception is the Key to Safety.......................................................4-74 The Output of the Track Circuit ......................................................................................4-77 Shunting .........................................................................................................................4-78 Remedies for Poor Shunting ..........................................................................................4-82 Railroad Track Circuits One at a Time ........................................................................4-83 Steady Energy DC Track Circuits ..............................................................................4-83 Polarized DC Track Circuits.......................................................................................4-86 Coded DC Track Circuits ...........................................................................................4-87

xviii

Trakode......................................................................................................................4-89 PMTC (Pulse-Modulated Track Circuit) .....................................................................4-90 Electro Code ..............................................................................................................4-91 Electrified Electro Code .............................................................................................4-92 Microtrax ....................................................................................................................4-93 AC Track Circuits In General ..................................................................................4-94 Classic AC Track Circuits ..........................................................................................4-94 Double Element Vane Relays ....................................................................................4-96 Phase Shifting............................................................................................................4-97 Centrifugal AC Relays................................................................................................4-97 Single-Rail Vane-Relay AC Track Circuits.................................................................4-98 Double-Rail AC Vane-Relay Track Circuits and Impedance Bonds ........................4-100 All-AC Signaling Systems ........................................................................................4-103 Style C Track Circuits ..............................................................................................4-104 Audio-Frequency Track Circuits...............................................................................4-105 Audio-Frequency Overlay ........................................................................................4-107 Impulse Track Circuits .............................................................................................4-109 Cab Signals .............................................................................................................4-109 Audio-Frequency Cab Signaling............................................................................4-112 Grade Crossings ......................................................................................................4-113 Crossing Predictors.............................................................................................4-114 Motion Sensors ...................................................................................................4-117 Motion Sensors, Crossing Predictors, and AC Interference .........................................4-118 Conventional or Stick Grade Crossing Warning Systems ................................4-120 Highway Crossing Gate Mechanisms .................................................................4-121 Highway Crossing Flashers ................................................................................4-123 Crossing Bells .....................................................................................................4-124 Highway Crossing Traffic Preemption......................................................................4-125 Alternatives to Track Circuits........................................................................................4-126 Communication-Based Crossing Warning Systems ................................................4-126 Magnetometers ........................................................................................................4-127 Wheel Detectors ......................................................................................................4-127 Axle Counters ..........................................................................................................4-128 Global Positioning Systems .....................................................................................4-128

xix

Hazard Detectors..........................................................................................................4-130 Wheel Defect Detectors ...........................................................................................4-131 Hotbox Detectors .....................................................................................................4-131 Slide Fences ............................................................................................................4-131 Tag Readers ............................................................................................................4-132 End-of-Train Monitoring (E.O.T.M.) .........................................................................4-133 Other Miscellaneous Track-Mounted Signal Equipment ..............................................4-134 Gauge Plates and Gauge Rods ...............................................................................4-134 Switch Circuit Controllers.........................................................................................4-136 Electric Switch Locks ...............................................................................................4-137 Switch Machines ......................................................................................................4-138 Impedance Bonds ....................................................................................................4-141 Other Signaling Topics .................................................................................................4-146 Electric Traction Circuits ..........................................................................................4-146 Safety, Vitality, and Fail-Safe Design.......................................................................4-149 Rules, Standards and Instructions ...........................................................................4-153 Lightning Protection and Grounding ........................................................................4-154 AC Interference Issues ............................................................................................4-157 References ........................................................................................................................4-160 5 ABNORMAL OPERATION OF RAILROAD EQUIPMENT ....................................................5-1 Introduction ...........................................................................................................................5-2 Quick-Start Version ...............................................................................................................5-2 Detailed Version ....................................................................................................................5-2 Abnormal Operation of Railroad Signal Equipment..........................................................5-2 Definitions, Terminology, and Philosophy....................................................................5-3 Redundancy vs. Vitality ...............................................................................................5-5 Dual vs. Vital ................................................................................................................5-6 Abnormal Operation.....................................................................................................5-7 Wrong-Side Abnormal Operation..........................................................................5-7 Right-Side Abnormal Operation............................................................................5-9 Abnormal Operation from Spoofing of Normal Events........................................5-11 Philosophy .................................................................................................................5-12 Methods of Achieving Vitality .....................................................................................5-13 Vitality and Creative Equipment Maintenance.........................................................5-14

xx

Rules, Vitality, and Wrong-Side Failures ...................................................................5-15 Railroad Signal Equipment Standards vs. Other Industries.......................................5-16 6 DAMAGE TO RAILROAD EQUIPMENT................................................................................6-1 Introduction ...........................................................................................................................6-2 Quick-Start Version ...............................................................................................................6-2 Detailed Version ....................................................................................................................6-2 Standards .........................................................................................................................6-3 Railroad Signal Equipment Design Practices ...................................................................6-5 Exposures.........................................................................................................................6-7 Mitigation ..........................................................................................................................6-9 7 PERSONNEL SAFETY CONSIDERATIONS .........................................................................7-1 Introduction ...........................................................................................................................7-2 Quick-Start Version ...............................................................................................................7-2 Shock Hazard ...................................................................................................................7-2 Electric Induction .........................................................................................................7-2 Magnetic Induction.......................................................................................................7-2 Fault Induced Voltages.....................................................................................................7-3 Electric and Magnetic Field Exposure ..............................................................................7-3 Detailed Version ....................................................................................................................7-4 Shock Hazard ...................................................................................................................7-4 Background..................................................................................................................7-4 Duration of Exposure ...................................................................................................7-6 Steady-State Exposure ................................................................................................7-6 Short Duration Exposure..............................................................................................7-7 Philosophy of Setting Standards..................................................................................7-9 Usual Circumstances ............................................................................................7-11 Steady-State Limits....................................................................................................7-12 Body/Contact Impedance......................................................................................7-12 Conductor Circuit Impedance................................................................................7-14 Source Impedance ................................................................................................7-14 Voltage Limits Derived From Current Effects........................................................7-16 Other Considerations ............................................................................................7-16 Possible Target Values .........................................................................................7-18

xxi

Limits for Faults and other Short-Duration Events .....................................................7-20 Possible Target Values .........................................................................................7-23 Electric and Magnetic Field Standards and Guidelines ..................................................7-23 Introduction to Standards and Guidelines..................................................................7-23 State Standards and Recommendations Related to Transmission Lines..................7-23 Guidelines for Exposure to 50/60 Hz Electric and Magnetic Fields ...........................7-25 International Standards..............................................................................................7-27 Swedish Standards for Computers and Monitors ......................................................7-29 Safe Working Clearances from Power Lines..................................................................7-30 References ..........................................................................................................................7-32 8 FIELD MEASUREMENTS ......................................................................................................8-1 Introduction ...........................................................................................................................8-2 General Precautions for all Test Procedures ........................................................................8-2 Tests .....................................................................................................................................8-3 General Information...............................................................................................................8-4 Test #1 Excessive Common Mode Voltage...................................................................8-7 Objectives ....................................................................................................................8-7 Overview ......................................................................................................................8-7 Equipment Required ....................................................................................................8-7 Procedure ....................................................................................................................8-7 Interpretation of Measurements ...................................................................................8-8 First Discriminant: Rail-to-Ground Voltage Magnitude............................................8-8 Second Discriminant: Rail-to-Ground Voltage Magnitude.......................................8-8 Test #2 Measure Dominant Frequency .......................................................................8-10 Objectives ..................................................................................................................8-10 Overview ....................................................................................................................8-10 Equipment Required ..................................................................................................8-10 Procedure ..................................................................................................................8-10 Interpretation of Measurements .................................................................................8-12 First Discriminant: Rail-to-Ground Voltage Magnitude..........................................8-12 Second Discriminant: Dominant Frequency..........................................................8-12 Test #3 Check for Excessive Track Circuit Unbalance................................................8-13 Objectives ..................................................................................................................8-13 Overview ....................................................................................................................8-13

xxii

Equipment Required ..................................................................................................8-13 Procedure ..................................................................................................................8-13 Interpretation of Measurements .................................................................................8-15 First Discriminant: Degree of Track Circuit Unbalance .........................................8-15 Test #4 - Spectral Analysis Test.....................................................................................8-16 Objective....................................................................................................................8-16 Overview ....................................................................................................................8-16 Equipment Required ..................................................................................................8-16 Procedure ..................................................................................................................8-16 Interpretation..............................................................................................................8-17 First Discriminant: Ratio of AC Power Fundamental to First Odd Harmonic.........8-17 Second Discriminant: Fundamental to First Even Harmonic Ratio .......................8-18 Third Discriminant: Power and Proximity to Motion Sensor or Crossing Predictor Operating Frequency .............................................................................8-18 Test #5 Coarse Rail Balance Test...............................................................................8-20 Objectives ..................................................................................................................8-20 Overview ....................................................................................................................8-20 Equipment Required ..................................................................................................8-20 Procedure ..................................................................................................................8-20 Interpretation of Measurements .................................................................................8-22 First Discriminant: Rail-to-Rail Voltage Magnitude................................................8-22 Second Discriminant: Rail-to-Ground Voltage Magnitude.....................................8-22 Third Discriminant: Rail-to-Ground Voltage Magnitude.........................................8-22 Fourth Discriminant: Rail-to-Ground Voltage Balance ..........................................8-22 Fifth Discriminant: Voltage Summation .................................................................8-23 Test #6 Rail-Rail and Rail-Ground Spatial Voltage Distribution Test ..........................8-24 Objective....................................................................................................................8-24 Overview ....................................................................................................................8-24 Equipment Required ..................................................................................................8-24 Procedure ..................................................................................................................8-24 Interpretation..............................................................................................................8-26 First Discriminant: Rail-to-Ground Voltage............................................................8-27 Second Discriminant: Rail-to-Rail Voltage ............................................................8-27 Third Discriminant: Rail-to-Ground Voltage Anomalies.........................................8-27 Test #7 Insulated Joint Test ........................................................................................8-28

xxiii

Objective....................................................................................................................8-28 Overview ....................................................................................................................8-28 Equipment Required ..................................................................................................8-29 Procedure ..................................................................................................................8-30 Interpretation..............................................................................................................8-32 Test #8 Direct Measurement of Insulated Joint Resistance ........................................8-34 Objective....................................................................................................................8-34 Overview ....................................................................................................................8-34 Equipment Required ..................................................................................................8-34 Procedure ..................................................................................................................8-34 Interpretation..............................................................................................................8-35 Test #9 Alternate Insulated Joint Test No Special Equipment .................................8-38 Objective....................................................................................................................8-38 Overview ....................................................................................................................8-38 Equipment Required ..................................................................................................8-38 Procedure ..................................................................................................................8-38 Interpretation..............................................................................................................8-39 Test #10 Arrester/Equalizer Test.................................................................................8-40 Objective....................................................................................................................8-40 Overview ....................................................................................................................8-40 Equipment Required ..................................................................................................8-40 Procedure ..................................................................................................................8-41 Interpretation..............................................................................................................8-42 Test #11 Hardwire Shunt Testing ................................................................................8-43 Objective....................................................................................................................8-43 Overview ....................................................................................................................8-43 Equipment Required ..................................................................................................8-43 Procedure ..................................................................................................................8-44 Interpretation..............................................................................................................8-44 Test #12 Local Ground/Power Company Neutral Isolation Test .................................8-45 Objective....................................................................................................................8-45 Overview ....................................................................................................................8-45 Equipment Required ..................................................................................................8-45 Procedure ..................................................................................................................8-45

xxiv

Interpretation..............................................................................................................8-46 9 INVESTIGATION ....................................................................................................................9-1 Introduction ...........................................................................................................................9-2 Quick-Start Version ...............................................................................................................9-2 Troubleshooting Fundamentals The Scientific Method .................................................9-2 Step 1): Recognition A Report of a Problem .................................................................9-2 Step 2): Study the Problem ..............................................................................................9-3 Step 3): Guess at the Cause(s) ........................................................................................9-3 Step 4): Predict the Implications.......................................................................................9-4 Step 5): Test Each Implication..........................................................................................9-4 Step 6): Form the Simplest Theory that Explains the Results ..........................................9-4 Is it Really AC Interference?.............................................................................................9-4 The Rules of Thumb of Railroad Signals and AC Interference ......................................9-5 Detailed Version ....................................................................................................................9-5 Introduction.......................................................................................................................9-5 The Nature of the Problem ...............................................................................................9-6 Troubleshooting Fundamentals The Scientific Method .................................................9-7 Recognizing AC Interference Problems.......................................................................9-7 Step 1): Recognition A Report of a Problem .................................................................9-9 Step 2): Study the Problem ............................................................................................9-10 Step 3): Guess at the Cause(s) ......................................................................................9-11 Step 4): Predict the Implications.....................................................................................9-12 Step 5): Test Each Implication........................................................................................9-13 Step 6): Form the Simplest Theory that Explains the Results. .......................................9-14 A Hidden Fork in the Road .............................................................................................9-15 The Source Path Receptor Model ............................................................................9-16 The Rules of Thumb of Railroad Signals and AC Interference ....................................9-17 The Questions ................................................................................................................9-20 The What Questions................................................................................................9-20 The Where Questions..............................................................................................9-21 The When Questions...............................................................................................9-22 Guessing at the Causes.............................................................................................9-23 The Specific Questions ..............................................................................................9-24 Common Sources for E.M. Fields Power Lines ...........................................................9-24

xxv

Common Sources for Conducted Interference ...............................................................9-25 Common Paths Fields, Rails, and Wires .....................................................................9-25 Common Receptors Motion Sensors and Crossing Predictors ...................................9-27 Motion Sensors and Crossing Predictors Symptoms and What They Mean ..........9-27 Case #1: The Crossing Equipment is Activated Continuously or Intermittently When there are No Trains Around ........................................................................9-28 Case #2: The Crossing Equipment is Activated Before it should be, but there is a Train Approaching ..........................................................................................9-31 Case #3: The Crossing Equipment Stays Activated for Longer than it should after a Train has Passed Completely through the Grade Crossing.......................9-33 Case #4: The Crossing Equipment is Activated Briefly after a Train has Passed Completely Through the Grade Crossing and the Crossing has Recovered.............................................................................................................9-33 10 DIAGNOSTIC FLOW CHART ............................................................................................10-1 Introduction .........................................................................................................................10-2 Following the Flow Chart.....................................................................................................10-3 Where to begin ...............................................................................................................10-3 How to Proceed ..............................................................................................................10-3 11 GRAPHIC EVALUATION OF PROPOSED CHANGES BASED ON SIMULATIONS .......11-1 Introduction .........................................................................................................................11-2 Summary of Graphical Aids for Compatibility Issues ..........................................................11-2 Personnel Safety for Steady State Magnetic Field Induction to Rails ............................11-7 Signal System Equipment Compatibility for Steady State Magnetic Field Induction to Rails..........................................................................................................................11-10 Personnel Safety for Steady State Electric Induction to Open Wire Signal Pole Lines .............................................................................................................................11-13 Personnel Safety for Faulted Power-Line Magnetic Field Induction to Rails ................11-13 Power-Line Fault Induced Rail Current for Lightning Arrester Survival ........................11-14 Example of Preliminary Compatibility Assessment by Graphical Procedures...................11-15 Introduction and Background........................................................................................11-15 Preliminary Review of Coupled Voltage and Current...............................................11-16 Steady State Coupling to Track ..........................................................................11-16 Steady State Coupling to Pole Line ....................................................................11-19 Fault Current Coupling to Track ..........................................................................11-19 Conclusions .............................................................................................................11-20

xxvi

12 COMPUTER MODELING OF JOINT CORRIDORS...........................................................12-1 Introduction .........................................................................................................................12-2 Quick-Start Version .............................................................................................................12-2 What Computer Modeling Will and Will Not Do ..............................................................12-2 General...........................................................................................................................12-2 Required Information ......................................................................................................12-3 Results............................................................................................................................12-7 Choosing a Consultant ...................................................................................................12-8 Choosing Software .........................................................................................................12-8 Detailed Version ..................................................................................................................12-8 Overview of Power System-Railroad Interactions ..........................................................12-8 Power Systems Steady State Magnetic Field for Railroad Compatibility Studies...........12-9 Quantifying Magnetic Field Coupling .........................................................................12-9 Three-Phase Systems .............................................................................................12-14 Unbalance Comparison Example ............................................................................12-19 More on Power Line Unbalance...............................................................................12-20 Conservative Effect of Steady State Unbalance Example ......................................12-21 Double Circuit Arrangements...................................................................................12-22 Steady State Compatibility Modeling of Railroad Systems for Magnetic Field Coupling from Power Systems .....................................................................................12-25 Track Signal Circuits and Their Effect on Induced Voltage......................................12-25 Track Unbalances and Induced Voltage at Signal Equipment .....................................12-34 Degraded IJ and Track Unbalance ..........................................................................12-42 Shorted IJ Example.............................................................................................12-46 Field Measurements............................................................................................12-47 Analysis of Exposure...........................................................................................12-51 Model Comparison to Field Measurements ........................................................12-51 Power-System Phasing.......................................................................................12-54 Degraded IJ with Train.............................................................................................12-55 Field-Measured Example ....................................................................................12-56 Simple Model with Shorted Insulator...................................................................12-58 Signal Equipment Impedance ..................................................................................12-61 Shorted Arrester to Ground......................................................................................12-63 Affect of Signal-System Ground Resistance on Coupled Interference ...............12-64 Conduction Coupling to Track Circuit..................................................................12-67

xxvii

Analysis Procedure for Evaluating Induced Voltage at Track Signal Equipment.....12-69 Susceptibility of Track Signal Equipment.................................................................12-71 Steady State Coupling to Railroad Signal Lines...........................................................12-71 Magnetic Field Coupling to Signal Lines..................................................................12-71 Electric Induction Coupling to Signal Lines Concepts...........................................12-72 Electric Induction Coupling to Signal Lines Example............................................12-76 Power Line Fault Coupling to Railroad System ............................................................12-83 Overview ..................................................................................................................12-83 Fault-Induced Firing of Track Arresters ...................................................................12-84 The Firing of Arresters in One Track Circuit........................................................12-85 The Firing of Arresters in Adjacent Track Circuits...............................................12-86 Fault-Induced Rail Current.......................................................................................12-90 Fault-Induced Rail Voltage and Personnel Safety ...................................................12-95 References ......................................................................................................................12-100 13 MITIGATION OPTIONS......................................................................................................13-1 Introduction .........................................................................................................................13-3 Mitigation Tables .................................................................................................................13-3 Detailed Options..................................................................................................................13-7 Power System Mitigation ................................................................................................13-7 Counterpoise Underground.....................................................................................13-7 Counterpoise Aerial ................................................................................................13-8 Fault Current Limiting.................................................................................................13-8 Load Current Limiting.................................................................................................13-9 Phase Arrangement Optimization (Cancellation).......................................................13-9 Split-Phasing............................................................................................................13-10 Neutral Wire Size Increase ......................................................................................13-10 Phase Current Balance Optimization.......................................................................13-11 Phase Spacing reduction (Phase Compaction) .......................................................13-12 Vertical Spacing Increase (Higher Structures).........................................................13-12 Horizontal Spacing Increase ....................................................................................13-13 Automatic Reclosing Restriction ..............................................................................13-13 Manual Reclosing Restriction ..................................................................................13-14 Fault Current Clearing Time Reduction ...................................................................13-14 Open Delta Transformer Removal ...........................................................................13-15

xxviii

Capacitor Bank Inspection/Monitoring .....................................................................13-15 Railroad Mitigation........................................................................................................13-16 Reducing Length of Electrically Continuous Track (At Power Frequencies) (Shorten Track Blocks) ............................................................................................13-16 Counterpoise Underground...................................................................................13-16 Grounding Un-Signaled Track/Sidings (Multigrounded Rail can Act as a Counterpoise) ..........................................................................................................13-17 Insulated Joint Inspection Increased.....................................................................13-18 Insulated Joint Replacement-Shorted or Leaky Units..............................................13-18 Surge Arresters Inspection Increased...................................................................13-19 Surge Arresters Replacement Failed Units...........................................................13-20 By-Pass Shunts And Couplers Removal .................................................................13-21 Resistive Network Installation ..................................................................................13-21 Isolation Transformer Installation.............................................................................13-22 Work Gloved (Work Using Electrically Insulated Rated Gloves.).............................13-22 Track Circuit Balance Optimization..........................................................................13-23 Battery Chargers Supplied at 240 Volts...................................................................13-23 Track Circuit Lead Fusing ........................................................................................13-24 Operating Frequencies Chosen to Maximize Immunity ...........................................13-25 Signal Output Level Increases .................................................................................13-26 Equipment Replaced with Less Susceptible Equivalent Systems ...........................13-27 Pole-Line Communication Systems Removed.........................................................13-27 Style C Track Circuits Removed............................................................................13-28 60 Hz Cab Signals Changed to Another Frequency ................................................13-29 Railroad Bed (Ballast) Condition Improvement........................................................13-29 Ground Mats/Grids Around Equipment....................................................................13-30 Safe Working Clearances from Power Lines................................................................13-31 14 CASE STUDIES .................................................................................................................14-1 Introduction .........................................................................................................................14-1 Excessive Common Mode Voltage .....................................................................................14-2 Case Study #1 ................................................................................................................14-2 Case Study #2 ................................................................................................................14-4 Case Study #3 ................................................................................................................14-6 Equipment Damage from Power System Events ................................................................14-7

xxix

Case Study #4 ................................................................................................................14-7 Case Study #5 ................................................................................................................14-8 Case Study #6 ................................................................................................................14-9 Case Study #7 ..............................................................................................................14-10 Steady-State Effects to Equipment Operation...................................................................14-11 Case Study #8 ..............................................................................................................14-11 Case Study #9 ..............................................................................................................14-12 Case Study #10 ............................................................................................................14-14 Case Study #11 ............................................................................................................14-15 Case Study #12 ............................................................................................................14-16 Case Study #13 ............................................................................................................14-17 15 PLANNING GUIDELINES FOR NEW/UPGRADED RAILROAD SYSTEMS.....................15-1 Introduction .........................................................................................................................15-2 Quick-Start Version .............................................................................................................15-2 Detailed Version ..................................................................................................................15-4 Sources ..........................................................................................................................15-4 Paths ..............................................................................................................................15-5 Distance.....................................................................................................................15-5 Block Length, Wideband Shunts, and Tuned Couplers .............................................15-6 Ballast ........................................................................................................................15-6 Receptors .......................................................................................................................15-7 Steady-State...................................................................................................................15-8 Fault................................................................................................................................15-9 General...........................................................................................................................15-9 Notification....................................................................................................................15-10 16 PLANNING GUIDELINES FOR NEW/UPGRADED POWER SYSTEMS ..........................16-1 Introduction .........................................................................................................................16-2 Quick-Start Version .............................................................................................................16-2 Detailed Version ..................................................................................................................16-4 Sources ..........................................................................................................................16-4 Voltage.......................................................................................................................16-4 Current Steady State ..............................................................................................16-5 Current Fault and Switching....................................................................................16-5

xxx

Phasing......................................................................................................................16-5 Balance......................................................................................................................16-5 Transmission .............................................................................................................16-6 Distribution.................................................................................................................16-6 Paths ..............................................................................................................................16-6 Magnetic Induction.....................................................................................................16-6 Electric Induction .......................................................................................................16-7 Earth Conduction .......................................................................................................16-7 Receptors .......................................................................................................................16-7 Steady-State...................................................................................................................16-7 Fault................................................................................................................................16-7 Notification......................................................................................................................16-7 17 CONCLUSIONS .................................................................................................................17-1 Current State of Affairs........................................................................................................17-2 Short-Term Solutions ..........................................................................................................17-2 Long-Term Solutions ...........................................................................................................17-3 Highway Grade Crossing Train Detection Equipment False Activation and Cab Signaling Equipment Operated at 60 Hz ........................................................................17-3 Damage to SPDs from Power System Fault Magnetic Induction ...................................17-4 Future Research .................................................................................................................17-5 Conclusion ..........................................................................................................................17-5 18 GLOSSARY........................................................................................................................18-1 Introduction .........................................................................................................................18-2 Glossary ..............................................................................................................................18-2 19 ERRATA SHEET ................................................................................................................19-1

xxxi

LIST OF FIGURES
Figure 2-1 The Source-Path-Receptor Interference Model........................................................2-5 Figure 2-2 Conductive Coupling Between Power and Railroad Systems ..................................2-7 Figure 2-3 Earth Conduction......................................................................................................2-7 Figure 2-4 Ground Potential Rise ..............................................................................................2-8 Figure 2-5 Wavelength Scaling when Measured in Wavelengths (), Printed Circuit Boards and Joint Railroad Power Corridors Present the Same Geometry ........................2-9 Figure 2-6 Capacitive Coupling to an Object Close to an Overhead Line ...............................2-10 Figure 2-7 Thevenin Equivalent Circuit for Electric Field Coupling ..........................................2-11 Figure 2-8 Norton Equivalent Circuit for Electric Field Coupling ..............................................2-11 Figure 2-9 Electric Field Induction ...........................................................................................2-11 Figure 2-10 Equivalent Circuit for Magnetic Field Coupling .....................................................2-12 Figure 2-11 Magnetic Field Induction.......................................................................................2-13 Figure 2-12 Shield Conductor for Magnetic Induction..............................................................2-14 Figure 2-13 Complete Circuit with Source, Path to the Load, Load, and Path from the Load .................................................................................................................................2-16 Figure 2-14 Circuit with Differential Mode Signal .....................................................................2-17 Figure 2-15 Circuit with Common Mode Signal .......................................................................2-18 Figure 2-16 Circuit with Common Mode Interference and Differential Mode Signal ................2-19 Figure 2-17 Common Mode Voltage Converted to Differential Mode ......................................2-20 Figure 2-18 Circuit with Different Resistances to Ground ........................................................2-21 Figure 2-19 Circuit with Unbalance: Common Mode Interference Converted into Differential Mode Interference..........................................................................................2-21 Figure 2-20 Differential Mode and Common Mode Voltages on Railroad Signals...................2-22 Figure 2-21 European Interference Model ...............................................................................2-23 Figure 2-22 North American Interference Model......................................................................2-24 Figure 2-23 FCC Part 15 Radiated Emissions Limits for Digital Devices.................................2-30 Figure 3-1 Overview of the Electric Power System....................................................................3-5 Figure 3-2 Overhead Transmission Structure Types .................................................................3-9 Figure 3-3 Double Circuit Line with 2 Shield Wires 8 Total Conductors ...............................3-11 Figure 3-4 Distribution Transformer Feeding Single-Phase 120/240V Load from a ThreePhase wye-Connected 12,470/7200V Primary with a Multi-Grounded Neutral...............3-14 Figure 3-5 Distribution Pole with Telephone and Cable Television Wires ...............................3-15

xxxiii

Figure 3-6 4-Wire Multi-Grounded Feeders .............................................................................3-16 Figure 3-7 3-Wire Ungrounded Feeders ..................................................................................3-18 Figure 3-8 Single Wire Earth Return Primary ..........................................................................3-21 Figure 3-9 Positive Sequence Currents Sum to Zero in a Balanced 3-Phase System ............3-22 Figure 3-10 Zero Sequence Harmonic Current Returns in the Neutral Path ...........................3-23 Figure 3-11 Pad Mounted Transformers ..................................................................................3-24 Figure 3-12 Pole Mounted Potheads for Overhead/Underground Transitions.........................3-25 Figure 4-1 Picture of Railroad Tracks ........................................................................................4-5 Figure 4-2 Typical AREMA Rail Section (133 lb. Rail) ...............................................................4-7 Figure 4-3 Thermite Weld in Rail ...............................................................................................4-7 Figure 4-4 Function and Designation of Rails............................................................................4-8 Figure 4-5 Guard Rails Minimize Damage Caused by Derailed Wheels ...................................4-9 Figure 4-6 Pictures of Railroad Ties, Wood and Concrete ......................................................4-10 Figure 4-7 Bridges Pose Special Track Circuit Problems, Due to their (Conductive) Steel Construction .....................................................................................................................4-11 Figure 4-8 Close-Up Pictures of a Tie-Plate, a Spike, and the Installed Assembly at the Foot of a Rail....................................................................................................................4-11 Figure 4-9 Picture of a Pigtail Rail Clip and Insulating Pad on a Concrete Tie......................4-12 Figure 4-10 Modern Track Construction Techniques can Provide Very High Ballast Resistance. (U.S.&S Inc.) Here we see Rails Resting on Concrete Ties that are Themselves Set in Concrete, Instead of Crushed Rock ..................................................4-14 Figure 4-11 Ordinary and Insulated Rail Joints........................................................................4-16 Figure 4-12 Welded Bondwires and Mechanically-Attached Bondwires..................................4-17 Figure 4-13 Jointed Rails with Railhead Cadweld Bonds, Long Bonds, and High-Current Bonds ...............................................................................................................................4-19 Figure 4-14 Unoccupied Track.................................................................................................4-21 Figure 4-15 Occupied Track ....................................................................................................4-22 Figure 4-16 Grade Crossing ....................................................................................................4-23 Figure 4-17 Impedance vs. Train Position ...............................................................................4-24 Figure 4-18 Typical Insulated Joint Installed in Rail, with Polyurethane-Encapsulated Joint Bars .........................................................................................................................4-26 Figure 4-19 Cross-Section View of a Typical Bonded Insulated Joint (Portec Rail Products, Inc.) ..................................................................................................................4-28 Figure 4-20 Sectional View of Rail and Typical Encapsulated Insulated Joint (Portec Rail Products, Inc.)...........................................................................................................4-28 Figure 4-21 Metallic Dust and Slivers Tend to Accumulate, Bridging Rails at Insulated Joint..................................................................................................................................4-30 Figure 4-22 Robinsons 1872 Track Circuit Patent ..................................................................4-32 Figure 4-23 Basic DC Track Circuit .........................................................................................4-33 Figure 4-24 Basic Track Circuit Schematic ..............................................................................4-34 Figure 4-25 Occupied DC Track Circuit ...................................................................................4-35

xxxiv

Figure 4-26 Relay Armature and Coil, Safetran Systems Type ST Plug-In Rely (w/o cover) ...............................................................................................................................4-36 Figure 4-27 Track Circuit Showing Broken Rail .......................................................................4-36 Figure 4-28 Rear View of a Wayside Signal Head (Searchlight Type) ....................................4-42 Figure 4-29 Section vs. Block ..................................................................................................4-44 Figure 4-30 Limited Clearance for Signals and other Equipment on a Rapid Transit Line. .....4-45 Figure 4-31 Elementary Interlocking Layouts ..........................................................................4-46 Figure 4-32 Mechanical Interlocking Machine .........................................................................4-47 Figure 4-33 Typical Power Interlocking Machine .....................................................................4-48 Figure 4-34 Rack-Mounted Plug-In Vital Relays in a Relay-Based Interlocking Plant .............4-49 Figure 4-35 Sophisticated Electronics Typical of Solid-State Interlocking Systems ................4-49 Figure 4-36 Common Types of Light Signals...........................................................................4-51 Figure 4-37 Track Plan Symbology for Wayside Signals .........................................................4-51 Figure 4-38 3-Aspect and 4-Aspect Signal Systems ...............................................................4-52 Figure 4-39 Colorlight signals (Safetran Systems) ..................................................................4-52 Figure 4-40 Colorlight Signals..................................................................................................4-53 Figure 4-41 Flashing Lights from a Grade Crossing (Shown Without Visor or Background) (Safetran Systems) .....................................................................................4-54 Figure 4-42 LED Wayside Signal Lamps (Safetran Systems) .................................................4-54 Figure 4-43 Rear View of a Wayside Signal Head (Searchlight Type) ....................................4-55 Figure 4-44 Safetran Unilens Signal (Safetran Systems)......................................................4-56 Figure 4-45 Sequence of Aspects for Typical Single Direction Automatic Block System (ABS) in Multi-Track Territory..........................................................................................4-60 Figure 4-46 Bi-directional Absolute-Permissive Block System (APB)......................................4-60 Figure 4-47 Overlap Versus Non-Overlapping Block Systems ................................................4-61 Figure 4-48 #20 Switches Allow Crossover Moves Between Tracks at 45 mph in CTC Territory. This Greatly Enhances the Speed and Efficiency of Train Movements ...........4-62 Figure 4-49 Typical CTC System.............................................................................................4-62 Figure 4-50 Complex Interlocking Control Panel on Heavy Rail Rapid Transit System...........4-63 Figure 4-51 Computer-Based Operations Control Center for a Light-Rail Transit System (U.S.&S Inc.) ....................................................................................................................4-63 Figure 4-52 Mechanical Trip Stop, Wayside Element ..............................................................4-64 Figure 4-53 Mechanical Trip Stop, Carborne Element.............................................................4-65 Figure 4-54 Magnetic Trip Stop System ..................................................................................4-66 Figure 4-55 Modern Driverless Automatic Train Control Systems Promote Efficient Land Utilization in Densely Populated Urban Area (U.S.&S Inc.) ............................................4-67 Figure 4-56 Primary Railroad Signal Batteries (Lalande Cells). New (Left), and Discharged (Right) ...........................................................................................................4-70 Figure 4-57 Modern Single Lead-Acid (or Ni-Cd) Battery Cell .................................................4-71 Figure 4-58 Simple Schematic Drawing of Cell, Resistor, and Battery Choke.........................4-72

xxxv

Figure 4-59 Drawing of Typical Track Circuit Battery Choke (Safetran Systems Corp.)..........4-72 Figure 4-60 Transmitter Coupling Transformer and Capacitor ................................................4-73 Figure 4-61 Typical Instrument Case Containing Shelf Mounted Vital Relays.........................4-74 Figure 4-62 Typical Instrument Case Containing Plug-in Vital Relays.....................................4-75 Figure 4-63 Shunting of the Track Circuit ................................................................................4-79 Figure 4-64 Shunting Efficiency of Track Circuit......................................................................4-80 Figure 4-65 Elementary Steady Energy DC Track Circuit .......................................................4-84 Figure 4-66 Center-Fed DC Track Circuit ................................................................................4-85 Figure 4-67 Polarized DC Track Circuits for Transmission of Signal Control Information via the Rails......................................................................................................................4-87 Figure 4-68 Elementary Coded DC Track Circuit ....................................................................4-88 Figure 4-69 Typical Electronic Track Circuit ............................................................................4-91 Figure 4-70 Pulse Timing of Electro Code II, 3, and 4 .............................................................4-92 Figure 4-71 Microtrax Modulation Pattern................................................................................4-93 Figure 4-72 Vane Relay (Safetran Systems) ...........................................................................4-96 Figure 4-73 Single Rail Track Circuit .......................................................................................4-99 Figure 4-74 Single-Rail and Double-Rail Track Circuits ........................................................4-101 Figure 4-75 Double-Rail AC Vane-Relay Track Circuits ........................................................4-102 Figure 4-76 Diagram of Impedance Bonds and Insulated Joints ...........................................4-102 Figure 4-77 Large Shelf Relays Characteristic of an All-AC Signaling System .....................4-104 Figure 4-78 Impedance Bonds...............................................................................................4-107 Figure 4-79 Cab Signaling and Automatic Train Control were Well-Established Technologies by 1940 ....................................................................................................4-110 Figure 4-80 GCP Block Diagram ...........................................................................................4-115 Figure 4-81 Track a Island, Showing Track Wires .................................................................4-117 Figure 4-82 Typical Half-Barrier Highway Crossing Gate Mechanism...................................4-123 Figure 4-83 Schematic of Lights and Shunts .........................................................................4-124 Figure 4-84 Typical Wayside Hazard Detector Location having Hotbox and Hot Wheel Detectors........................................................................................................................4-131 Figure 4-85 Rock Slide in an Area Protected by an Overhead Slide Fence (Note the many insulators on the pole arms.) ................................................................................4-132 Figure 4-86 Typical End-of-Train-Monitor (E.O.T.M.) (U.S.&S. Inc.) .....................................4-134 Figure 4-87 Gauge Plates and Gauge Rods..........................................................................4-135 Figure 4-88 Insulation of Switch Gage Plates and Rods Prevents Unintended Short Circuiting of Track Circuits (U.S.&S. Inc.) ......................................................................4-135 Figure 4-89 Deteriorated Insulation and Reduced Clearances on Gauge Plates ..................4-136 Figure 4-90 Typical Hand-Operated Switch with Circuit Controller (silver box in upper right corner)....................................................................................................................4-137 Figure 4-91 Hand-Operated Switch Machine.........................................................................4-138 Figure 4-92 Typical Electric Yard Switch Machine with Circuit Controller (U.S.&S. Inc.).......4-139

xxxvi

Figure 4-93 Typical Dual-Control Electric Switch Machine ....................................................4-139 Figure 4-94 Typical Electro-Pneumatic Switch Machine and Valve.......................................4-140 Figure 4-95 Electro-Hydraulic Switch Operator (RTI, Inc.) ....................................................4-140 Figure 4-96 Typical Construction of Impedance Bond Used on Power Frequency Track Circuits in DC Propulsion Territory ................................................................................4-142 Figure 4-97 Typical Impedance Bond Layout, 750 VDC Propulsion Territory, 100 Hz Track Circuits; Impedance Bonds Rated 2500 Amps Per Rail.......................................4-143 Figure 4-98 Various Types of Resonating Circuits for Power Frequency Impedance Bonds .............................................................................................................................4-144 Figure 4-99 Typical Impedance Bond Layout, 12 kV, 25 Hz Propulsion Territory, 100 Hz Track Circuits; Impedance Bonds Rated 300 Amps per Rail .........................................4-145 Figure 4-100 Typical Impedance Bond Layout for Audio-Frequency Track Circuits, 750 VDC Propulsion Territory, 1-5 kHz Track Circuits; Impedance Bond Rated 3000 Amps per Rail.................................................................................................................4-145 Figure 4-101 Electrolysis of Underground Structures Caused by Unwanted DC Current Flow................................................................................................................................4-147 Figure 4-102 Electric Traction Fed by Overhead Wire and Contact Rail ...............................4-148 Figure 4-103 Impedance Bonds with Electric Traction ..........................................................4-149 Figure 4-104 Various Wayside Signal Mountings ..................................................................4-149 Figure 4-105 Track Circuits Provide Broken Rail Detection Capability. However, as Shown by this Cracked Joint Bar Bridged with a Bondwire, Certain Defects are Undetectable in the Track Circuit ...................................................................................4-151 Figure 4-106 Although Signal Systems and Track Circuits are Based Upon Fail-safe Design Principles, No System Offers Protection Against All Conceivable Hazards.......4-152 Figure 4-107 Typical Lightning Arrestors. Red Arrestors are Line-to-Ground, Blue Arrestors are Line-to-Line .............................................................................................4-155 Figure 4-108 Entrance Terminals with Wide, Flat Grounding Strips ......................................4-156 Figure 4-109 Effect of Multiple Grounds on Track Circuits ....................................................4-156 Figure 4-110 Adequate Clearance Between Signal Equipment and Overhead Lines Must be Maintained.................................................................................................................4-158 Figure 4-111 Foreign Energy on Track Circuits or other Signal Facilities may appear when Electric Switch Heaters Become Grounded to the Running Rails ........................4-159 Figure 5-1 Elementary Steady Energy DC Track Circuit ...........................................................5-4 Figure 5-2 Abnormal Operation: Classic vs. Vital Redundant Systems ................................5-6 Figure 6-1 Track Circuit Coupling Network ................................................................................6-7 Figure 7-1 Time/Current Zones of Effects of AC Currents 15 Hz to 100 Hz (refer to Table 7-2 for instructions)(IEC 479-1: Figure 14, 1994)...............................................................7-8 Figure 7-2 IEEE Std 80-1986 Guideline and IEC 479-1, 1994 - Survival Body Current ............7-9 Figure 7-3 Values of Total Body Impedance Hand-Hand or Hand-Foot for AC 50/60 Hz (Trend Lines Through Data Values Presented in IEC 479-1)...........................................7-13 Figure 7-4 Electrostatic Coupling to Conductor Touched by Man ...........................................7-15 Figure 7-5 IEEE-80 Fault Touch Potential Model ....................................................................7-21

xxxvii

Figure 7-6 IEEE Std 80-1986 Guideline for 99.5% Safe Touch Potential (70 Kg Person).......7-22 Figure 7-7 IEEE Std 80-1986 Guideline for 99.5% Safe Touch Potential (50 Kg Person).......7-22 Figure 7-8 Sample of a Working Clearances Card ..................................................................7-31 Figure 7-9 Railroad Signal Equipment with Marginal Working Clearance ...............................7-31 Figure 8-1 Magnetically Induced Rail Voltage ...........................................................................8-5 Figure 8-2 Effects of Track Circuit Unbalance on Magnetically Induced Voltage ......................8-6 Figure 8-3 Hz Button on Fluke 87 to Display Dominant Frequency of AC Voltage ...............8-11 Figure 8-4 Hand Held Audio Frequency Spectrum Analyzer ...................................................8-12 Figure 8-5 Spectrum of Rail-to-Ground Voltage Caused by AC Distribution Line....................8-18 Figure 8-6 Magnetically Induced AC Rail Voltages..................................................................8-26 Figure 8-7 Degraded IJ Converts Common Mode Interference into Differential Mode ............8-28 Figure 8-8 Insulated Joint Tester .............................................................................................8-31 Figure 8-9 Use of Insulated Joint Tester..................................................................................8-32 Figure 8-10 Photograph of IJ Leakage Current Test ...............................................................8-35 Figure 8-11 IJ Resistance from Voltage and Current...............................................................8-36 Figure 8-12 IJ Resistance from Voltage and Current Adjusted for Tested Current Probe ....8-37 Figure 8-13 Arrester Testing ....................................................................................................8-42 Figure 10-1 Railroad AC Power Interference Flow Chart ........................................................10-4 Figure 11-1 Vertical Power Line Configuration ........................................................................11-4 Figure 11-2 Horizontal Power Line Configuration ....................................................................11-5 Figure 11-3 Delta Power Line Configuration............................................................................11-6 Figure 11-4 Residual Current from Cramer Data .....................................................................11-7 Figure 11-5 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m ...............11-21 Figure 11-6 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m.................11-21 Figure 11-7 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m .............11-22 Figure 11-8 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m...............11-22 Figure 11-9 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m ...........11-23 Figure 11-10 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m .....11-23 Figure 11-11 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m.............11-24 Figure 11-12 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m ..............11-24 Figure 11-13 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m...........11-25

xxxviii

Figure 11-14 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m ............11-25 Figure 11-15 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m.........11-26 Figure 11-16 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m ..........11-26 Figure 11-17 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m........11-27 Figure 11-18 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m .........11-27 Figure 11-19 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m......11-28 Figure 11-20 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m .......11-28 Figure 11-21 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m....11-29 Figure 11-22 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m .....11-29 Figure 11-23 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m.............11-30 Figure 11-24 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m ..............11-30 Figure 11-25 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m...........11-31 Figure 11-26 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m ............11-31 Figure 11-27 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m.........11-32 Figure 11-28 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m ..........11-32 Figure 11-29 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m ............................................................................................................................11-33 Figure 11-30 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m ............................................................................................................................11-33 Figure 11-31 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m ............................................................................................................................11-34 Figure 11-32 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m ............................................................................................................................11-34

xxxix

Figure 11-33 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m ............................................................................................................................11-35 Figure 11-34 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m ............................................................................................................................11-35 Figure 11-35 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohmm ....................................................................................................................................11-36 Figure 11-36 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m ...11-36 Figure 11-37 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohmm ....................................................................................................................................11-37 Figure 11-38 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohmm ....................................................................................................................................11-37 Figure 11-39 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m ............................................................................................................................11-38 Figure 11-40 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohmm ....................................................................................................................................11-38 Figure 11-41 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks.....................................................................................................11-39 Figure 11-42 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks.....................................................................................................11-39 Figure 11-43 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks.....................................................................................................11-40 Figure 11-44 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks.....................................................................................................11-40 Figure 11-45 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks.....................................................................................................11-41 Figure 11-46 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks.....................................................................................................11-41 Figure 11-47 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohmm, 1 Mile Blocks .............................................................................................................11-42

xl

Figure 11-48 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks ..................................................................................................................11-42 Figure 11-49 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohmm, 1 Mile Blocks .............................................................................................................11-43 Figure 11-50 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohmm, 1 Mile Blocks .............................................................................................................11-43 Figure 11-51 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks.....................................................................................................11-44 Figure 11-52 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohmm, 1 Mile Blocks .............................................................................................................11-44 Figure 11-53 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks...............................................................................11-45 Figure 11-54 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks...............................................................................11-45 Figure 11-55 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks.............................................................................11-46 Figure 11-56 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks.............................................................................11-46 Figure 11-57 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks..........................................................................11-47 Figure 11-58 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks..........................................................................11-47 Figure 11-59 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks...............................................................................11-48 Figure 11-60 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks ................................................................................................11-48 Figure 11-61 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks.............................................................................11-49 Figure 11-62 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks ..............................................................................................11-49

xli

Figure 11-63 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks...........................................................................11-50 Figure 11-64 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks ............................................................................................11-50 Figure 11-65 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks...............................................................................11-51 Figure 11-66 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks...............................................................................11-51 Figure 11-67 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks.............................................................................11-52 Figure 11-68 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks.............................................................................11-52 Figure 11-69 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks..........................................................................11-53 Figure 11-70 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks..........................................................................11-53 Figure 11-71 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks...............................................................................11-54 Figure 11-72 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks...............................................................................11-54 Figure 11-73 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks............................................................................11-55 Figure 11-74 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks............................................................................11-55 Figure 11-75 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks..........................................................................11-56 Figure 11-76 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks..........................................................................11-56 Figure 11-77 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks.....................................................................................................11-57

xlii

Figure 11-78 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks.....................................................................................................11-57 Figure 11-79 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.....................................................................................................11-58 Figure 11-80 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.....................................................................................................11-58 Figure 11-81 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks.....................................................................................................11-59 Figure 11-82 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks.....................................................................................................11-59 Figure 11-83 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohmm, 2 Mile Blocks .............................................................................................................11-60 Figure 11-84 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks ..................................................................................................................11-60 Figure 11-85 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohmm, 2 Mile Blocks .............................................................................................................11-61 Figure 11-86 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohmm, 2 Mile Blocks .............................................................................................................11-61 Figure 11-87 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks.....................................................................................................11-62 Figure 11-88 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohmm, 2 Mile Blocks .............................................................................................................11-62 Figure 11-89 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks...............................................................................11-63 Figure 11-90 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks...............................................................................11-63 Figure 11-91 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.............................................................................11-64 Figure 11-92 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.............................................................................11-64

xliii

Figure 11-93 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks..........................................................................11-65 Figure 11-94 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks..........................................................................11-65 Figure 11-95 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks...............................................................................11-66 Figure 11-96 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks ................................................................................................11-66 Figure 11-97 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.............................................................................11-67 Figure 11-98 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks ..............................................................................................11-67 Figure 11-99 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks...........................................................................11-68 Figure 11-100 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks ............................................................................................11-68 Figure 11-101 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks..............................................................................11-69 Figure 11-102 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks..............................................................................11-69 Figure 11-103 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks............................................................................11-70 Figure 11-104 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.............................................................................11-70 Figure 11-105 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks...........................................................................11-71 Figure 11-106 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks...........................................................................11-71 Figure 11-107 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks...............................................................................11-72

xliv

Figure 11-108 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks...............................................................................11-72 Figure 11-109 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.............................................................................11-73 Figure 11-110 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks.............................................................................11-73 Figure 11-111 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks..........................................................................11-74 Figure 11-112 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks..........................................................................11-74 Figure 11-113 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Negative Side, Without Shield Wire ..........................11-75 Figure 11-114 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Positive Side, Without Shield Wire............................11-75 Figure 11-115 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Negative Side, With Shield Wire ...............................11-76 Figure 11-116 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Positive Side, With Shield Wire.................................11-76 Figure 11-117 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Negative Side, Without Shield Wire......................11-77 Figure 11-118 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Positive Side, Without Shield Wire .......................11-77 Figure 11-119 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Negative Side, With Shield Wire...........................11-78 Figure 11-120 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Positive Side, With Shield Wire ............................11-78 Figure 11-121 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Negative Side, Without Shield Wire ............11-79 Figure 11-122 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Positive Side, Without Shield Wire ..............11-79 Figure 11-123 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Negative Side, With Shield Wire .................11-80 Figure 11-124 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Positive Side, With Shield Wire ...................11-80 Figure 11-125 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 155-lb Person, Earth Resistivity 10 ohm-m ........................................11-81 Figure 11-126 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 110-lb Person, Earth Resistivity 10 ohm-m ........................................11-81

xlv

Figure 11-127 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 155-lb Person, Earth Resistivity 100 ohm-m ......................................11-82 Figure 11-128 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 110-lb Person, Earth Resistivity 100 ohm-m ......................................11-82 Figure 11-129 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 155-lb Person, Earth Resistivity 1000 ohm-m ....................................11-83 Figure 11-130 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 110-lb Person, Earth Resistivity 1000 ohm-m ....................................11-83 Figure 11-131 Calculated Single Phase to Ground Fault-Induced Rail Current for Long Exposure ........................................................................................................................11-84 Figure 11-132 Fault-Induced Rail Current Vs. Exposure Length ...........................................11-84 Figure 12-1 Electromagnetic Compatibility Information Form, Page 1 of 4..............................12-4 Figure 12-2 Electromagnetic Compatibility Information Form, Page 2 of 4..............................12-5 Figure 12-3 Electromagnetic Compatibility Information Form, Page 3 of 4..............................12-6 Figure 12-4 Electromagnetic Compatibility Information Form, Page 4 of 4..............................12-7 Figure 12-5 Geometry for Conductor and Observation Point Above the Earth......................12-11 Figure 12-6 Complex Image Plane Geometry .......................................................................12-13 Figure 12-7 Approximate Image Depth for Earth Return Conductor......................................12-14 Figure 12-8 Distances from Power Conductors to Observation Point....................................12-15 Figure 12-9 Phasor Components of Total Longitudinal Electric Field (LEF) for Figure 128 .....................................................................................................................................12-16 Figure 12-10 Normalized Balanced Component of LEF for Single Circuit Horizontal Line with No Shield Wires ......................................................................................................12-17 Figure 12-11 Magnitude of Balanced Component of LEF for Single Circuit Vertical Circuit without Shield Wires.......................................................................................................12-18 Figure 12-12 Normalized Residual Component of LEF for Single Vertical Circuit with No Shield Wires ...................................................................................................................12-19 Figure 12-13 Measured Phase Current Deviation from the Circuit Mean ..............................12-21 Figure 12-14 Predicted Rail-to-Ground Voltage Including 3% Current Unbalance ................12-22 Figure 12-15 Summary of Vertical Double Circuit Phase Assignments .................................12-23 Figure 12-16 Normalized Balanced Component of LEF for Double Circuit Vertical Center Point Symmetric Line with No Shield Wires ...................................................................12-24 Figure 12-17 Normalized Balanced Component of LEF for Double Circuit Vertical Super Bundle with No Shield Wires.........................................................................................12-25 Figure 12-18 Longitudinal Electric Field Excitation of Rails and Induced Voltages ...............12-26 Figure 12-19 Equivalent Circuit of Incremental Segment of Rail ...........................................12-27 Figure 12-20 Normalized Voltages, End of Rail-to-Remote Earth .........................................12-29 Figure 12-21 Normalized Voltages, End of Rail-to-Remote Earth .........................................12-30 Figure 12-22 Normalized Voltages, End of Rail-to-Remote Earth .........................................12-30 Figure 12-23 Ballast Resistivity Test Arrangement ................................................................12-31 Figure 12-24 Ballast Measurement Data Sheet.....................................................................12-32

xlvi

Figure 12-25 Measured Ballast Resistivity at Talmadge, Ohio ..............................................12-33 Figure 12-26 Equivalent Circuit at Metal-Electrolyte Interface...............................................12-34 Figure 12-27 Measured Resistance of Rail Insulated Joints..................................................12-35 Figure 12-28 Calculated Rail-to-Ground Voltage at IJ Locations for All 70-ohm IJs .............12-36 Figure 12-29 Calculated Rail-to-Ground Voltage at IJ Locations for All 30-ohm IJs .............12-37 Figure 12-30 Calculated Rail-to-Ground Voltage at IJ Locations for All 10-ohm IJs .............12-38 Figure 12-31 Calculated Rail-to-Ground Voltage at IJ Locations for 10-ohm IJs ..................12-39 Figure 12-32 Calculated Rail-to-Ground Voltage at IJ Locations for 10-ohm IJs ..................12-39 Figure 12-33 Calculated Rail-to-Ground Voltage at IJ Locations for 70-ohm IJ Resistance at Location 0.0.............................................................................................12-40 Figure 12-34 Calculated Rail-to-Ground Voltage at IJ Locations for 30-ohm IJ Resistance at Location 0.0.............................................................................................12-41 Figure 12-35 Calculated Rail-to-Ground Voltage at IJ Locations for 10-ohm IJ Resistance at Location 0.0.............................................................................................12-41 Figure 12-36 Rail-to-Ground Voltage vs Resistance of IJs at Zero Reference .....................12-42 Figure 12-37 Rail-to-Ground Voltage Profile With One Shorted Insulated Joint ....................12-43 Figure 12-38 Conversion of Rail-Ground Voltage to Rail-Rail Voltage vs. Degraded IJ Resistance for 1.5 mile track circuits.............................................................................12-44 Figure 12-39 Conversion of Rail-Ground Voltage to Rail-Rail Voltage vs. Degraded IJ Resistance for One Mile Track Circuits .........................................................................12-45 Figure 12-40 Measured Voltages at the Three Signal Locations on the Westbound Track...12-47 Figure 12-41 Sketch Showing the Location and Labeling of the Voltage Measurements Relative to the Rails .......................................................................................................12-48 Figure 12-42 Flexible Current Probe Being used to Measure Rail Current............................12-50 Figure 12-43 Calculated Rail-to-Rail Voltages .......................................................................12-53 Figure 12-44 Calculated Rail Voltages for the Same Conditions as Figure 12-43, but Without a Degraded Insulator .......................................................................................12-53 Figure 12-45 Calculated Rail Voltages ..................................................................................12-54 Figure 12-46 Track Sketches.................................................................................................12-56 Figure 12-47 Oscillograph Recording of the Rail-to-Rail Voltage on Each Side of the Insulated Joint ................................................................................................................12-57 Figure 12-48 Calculated Induced 60 Hz Rail-to-Rail Voltage at Insulated Joint Locations for Selected Operational Conditions ..............................................................................12-59 Figure 12-49 Calculated Results for the Same Four Analyzed Conditions of Figure 1248, Except with the Ballast Resistivity Reduced to 3 OhmKft .......................................12-60 Figure 12-50 Affect of Equipment Impedance on Induced Rail-to-Rail Voltage.....................12-62 Figure 12-51 Rail-to-Rail Voltage with a Shorted Track Arrester, Normalized by NonPerturbed Rail-to-Ground Voltage ...............................................................................12-65 Figure 12-52 Calculated Rail-to-Rail Voltage Profile Over the Length of the Track Circuit....12-66 Figure 12-53 Sketch of Track Circuit for Coupling of Interference from Distribution Neutral Through Shorted Track Arrester ........................................................................12-68

xlvii

Figure 12-54 Secondary Neutral to Earth Voltage [75] ..........................................................12-69 Figure 12-55 Power Line Voltage Coupling to Pole-Mounted Signal Conductor by Electric-Field Coupling ...................................................................................................12-73 Figure 12-56 Lateral Profiles of the Electric Field at Ground Level for Lines of Flat, Equilateral Delta, and Vertical Configurations [84]........................................................12-74 Figure 12-57 Current to Person Touching Conductor that is Capacitively Coupled to Power Line .....................................................................................................................12-76 Figure 12-58 Exposure of Rail System to 138-kV Transmission Line Typical Cross Section ...........................................................................................................................12-77 Figure 12-59 Calculated Space Potential For 138-kV Transmission Line of Figure 12-58 ....12-78 Figure 12-60 Electric Induction Voltage on Isolated Wire and Current to 1500-Ohm Person versus Excited Length of Wire ...........................................................................12-79 Figure 12-61 Effect of Total Body Impedance on Shock Current from Electric Induction Excited Isolated Conductor ............................................................................................12-80 Figure 12-62 Predicted Current to Earth Through 1500 Ohm Person From Electric Induction Excited Pole Line Signal Circuit......................................................................12-81 Figure 12-63 Predicted Open Circuit Voltage of Electric Induction Excited Signal Pole Line Circuits ...................................................................................................................12-82 Figure 12-64 Fault-Induced Voltage Across Track Lightning Arresters .................................12-85 Figure 12-65 Fault-Induced Rail Voltage Profiles with and Without a Fired Arrester.............12-86 Figure 12-66 Fault-Induced Arrester Voltages .......................................................................12-87 Figure 12-67 Fault-Induced Sequence of Track Arrester Firing Along Corridor.....................12-89 Figure 12-68 Fault-Induced Current at Fired Arresters ..........................................................12-90 Figure 12-69 Calculated Fault-Induced Rail Current for Long Exposure ...............................12-92 Figure 12-70 Single-Phase to Ground Fault Current vs. Fault Location ................................12-93 Figure 12-71 Calculated Fault-Induced Rail Current for Several Fault Locations..................12-94 Figure 12-72 Calculated Fault-Induced Rail Current for Several Fault Locations..................12-94 Figure 12-73 Fault-Induced Touch-Potential Safety Guidelines (70 kg person) ....................12-96 Figure 12-74 Fault-Induced Touch-Potential Safety Guidelines (50 kg person) ....................12-97 Figure 12-75 Relation Between Safe Touch Potential and Induced Voltage vs. Uniform Soil Resistivity ................................................................................................................12-98 Figure 12-76 Calculated Fault-Induced Rail-to-Earth Voltage for Several Fault Locations....12-99 Figure 12-77 Calculated Fault-Induced Rail-to-Earth Voltage for Several Fault Locations..12-100 Figure 13-1 Sample of a Working Clearances Card ..............................................................13-31 Figure 13-2 Railroad Signal Equipment with Marginal Working Clearance ...........................13-32 Figure 14-1 Capacitor Bank ...................................................................................................14-15 Figure 14-2 Fused Disconnects (Closed on Left, Open or Blown on Right) ..........................14-16 Figure 14-3 Shorted Surge Arrester.......................................................................................14-18 Figure 15-1 Moving Track Alignment Away from the Source can Lead to Increased Interference if a Buried Counterpoise was Used to Mitigate Magnetic Induction .............15-5 Figure 15-2 Typical Frequency Spectrum Measured Rail-to-Rail ............................................15-7

xlviii

Figure 15-3 Electromagnetic Compatibility Information Form, Page 1 of 4............................15-11 Figure 15-4 Electromagnetic Compatibility Information Form, Page 2 of 4............................15-12 Figure 15-5 Electromagnetic Compatibility Information Form, Page 3 of 4............................15-13 Figure 15-6 Electromagnetic Compatibility Information Form, Page 4 of 4............................15-14 Figure 16-1 Electromagnetic Compatibility Information Form, Page 1 of 4..............................16-9 Figure 16-2 Electromagnetic Compatibility Information Form, Page 2 of 4............................16-10 Figure 16-3 Electromagnetic Compatibility Information Form, Page 3 of 4............................16-11 Figure 16-4 Electromagnetic Compatibility Information Form, Page 4 of 4............................16-12

xlix

LIST OF TABLES
Table 2-1 Coupling Mechanisms Between Power Lines and Railroads ....................................2-3 Table 2-2 Common Mode and Differential Mode Coupling........................................................2-4 Table 2-3 Essential EMC Tests ...............................................................................................2-25 Table 2-4 Performance (failure) Criteria ..................................................................................2-27 Table 2-5 Basic EMC Standards..............................................................................................2-28 Table 2-6 EMC Requirements Using Basic Standards ............................................................2-29 Table 3-1 Component Parts of the Power System.....................................................................3-3 Table 3-2 Sources of Current Entering Earth from the Power System ......................................3-4 Table 4-1 Signal Aspect (Color) Displayed for Polarized DC Track Circuits............................4-86 Table 6-1 Damage Table, Taken from AREMA C&S Manual Part 11.5.1..................................6-4 Table 7-1 Reactions to Various Levels of Current[33] ...............................................................7-5 Table 7-2 Time/Current Zones for AC 15 Hz to 100 Hz (IEC 479-1: Table 4, 1994) .................7-7 Table 7-3 State Regulations that Limit Field Strengths on Transmission Line Rights-ofWay ..................................................................................................................................7-24 Table 7-4 Summary of ICNIRP 50/60 Hz Exposure Guidelines...............................................7-26 Table 7-5 Summary of ACGIH 60 Hz Exposure Guidelines ....................................................7-26 Table 7-6 Summary of International Electric Field Standards kV/m......................................7-27 Table 7-7 Summary of International Magnetic Field Standards Gauss (rms) .......................7-28 Table 7-8 Comparison of MPR II and TCO Guidelines............................................................7-30 Table 8-1 Diagnostic Flow Chart Tests ......................................................................................8-3 Table 8-2 Resolving Track Circuit Unbalance............................................................................8-3 Table 8-3 Locating Conducted Sources.....................................................................................8-3 Table 8-4 Voltage versus dBV Equivalence...............................................................................8-4 Table 11-1 Personnel Safety for Steady State Magnetic Field Induction to Rails....................11-8 Table 11-2 Signal System Equipment Compatibility for Steady State Magnetic Field Induction to Rails............................................................................................................11-11 Table 11-3 Personnel Safety for Steady State Electric Induction to Open Wire Signal Pole Lines ......................................................................................................................11-13 Table 11-4 Personnel Safety for Faulted Power-line Magnetic Field Induction to Rails ........11-14 Table 11-5 Power-Line Fault Induced Rail Current for Lightning Arrester Survival ...............11-15 Table 11-6 AREMA Guidelines for Interference Tolerance of Selected Signal Equipment....11-18 Table 12-1 Westbound Track IJ Resistance Measurements (June 2000) .............................12-49

li

Table 12-2 Conduction Voltage Coupling Conversion of 60-Hz Voltage on Signal Ground into Rail-to-Rail Voltage by Shorted Track Arrester (15-ohm Signal Ground Resistance) ....................................................................................................................12-68 Table 12-3 Values of Field Strength at 20 meters (65.6 feet) from Three Configurations of 525 kV Transmission Lines ........................................................................................12-74 Table 12-4 Signal Pole-Line Circuits Used to Evaluate Available Electric Induction Coupled Shock Current..................................................................................................12-81 Table 13-1 Excessive Common Mode Voltage Mitigation Power System And Railroad.......13-4 Table 13-2 Equipment Damage from Power System Events Mitigation Power System and Railroad.....................................................................................................................13-5 Table 13-3 Steady-State Effects to Equipment Operation Mitigation Power System and Railroad............................................................................................................................13-6 Table 15-1 Changes that tend to Increase Ac Interference Levels (Bad Things) ....................15-3 Table 15-2 Changes that tend to Decrease Ac Interference Levels (Good Things) ................15-3 Table 16-1 Changes that tend to Increase AC Interference Levels (Bad Things)....................16-3 Table 16-2 Changes that tend to Decrease AC Interference Levels (Good Things)................16-3

lii

1
INTRODUCTION AND BACKGROUND
This chapter provides a broad introduction to the issues surrounding the electrical effects that power transmission and distribution can have on railroads, and describes the evolution of these issues since the late 19th century. The definition of ac power interference is discussed, as is the need for common terminology between the two industries. The chapter concludes with a discussion of the organization of the information within the Handbook. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA.

1-1

Introduction and Background

Background
The problem of alternating current (ac) power lines affecting railroad signal circuits is not a new one, having been present to some degree since the advent of commercial ac power transmission (New York, circa 1893). However, changes in railroad signaling technology and the widespread use of joint electric utility / railroad corridors have greatly increased the potential for interference. Prior to commercial ac power transmission, direct current (dc) power systems were in use at numerous locations around the United States. Initially, they were generally restricted to small generating systems located within a few miles of their customers, due to voltage drops caused by the resistance of the wires. To address this issue, a separate pair of wires was usually run all the way from the generating station to each customer. Thus, the skies above the streets of New York City had become a tangled web of wires by the late 1880s. One solution, conceived and promoted by Nicola Tesla and George Westinghouse, respectively, was to use alternating current. Unlike direct current, this form of electrical power could be adjusted up and down in voltage as needed, using only simple transformers. This intern allowed the voltage to be increased dramatically at the generating station for efficient long-distance transmission, and then transformed down to the necessary voltage(s) for each customer. Thus alternating current, especially in its three-phase form, has remained the standard for electrical power transmission throughout the world for more than one hundred years. But, even though it provides for efficient transmission and flexible distribution systems, it does have a few drawbacks. The most significant of these within the railroad signaling world is that of ac interference both conducted interference and induction. Any conductor carrying an alternating current creates time-varying electric and magnetic fields around it. And any other conductor nearby can turn those time-varying electric and magnetic fields into a separate (but related) alternating current. This is the principle of electromagnetic induction. Any power system under fault conditions can create earth currents that conductively couple into railroad systems. Power systems employing multigrounded neutral conductors can provide conductively coupled ac interference continually. Where ac power lines and railroad signal circuits are concerned, these effects become greatest with conductors that run parallel and in close proximity to each other, over a long distance. Unfortunately, this is exactly the situation that occurs when ac power lines are located along a railroad right-of-way. The very large currents (hundreds of amps) carried by the power lines can cause significant ac power to be magnetically induced onto the railroad rails. The very large voltages (thousands to hundreds of thousands of volts) can cause significant electric induction on pole line circuits and train cars. In theory, where trains are powered by diesel locomotives, the very small ac currents (a few milliamps to a few amps) generated by some railroad signaling systems could also cause some induction into the power lines and in fact this is true. However, the amount of current induced into the power lines by the railroad signaling system is insignificant.

1-2

Introduction and Background

But let us consider the usual case. The current induced by the power lines into the railroad rails can be many times larger than the currents used by the railroad signaling system. Thus, this current can interfere with proper operation of the signaling system. So the interference usually occurs in one direction only. This unseen path of induction, together with conducted interference, forms the heart of the compatibility problem between ac power lines and railroad signaling circuits. So historically, the use of alternating current for electrical power transmission and distribution removed the need to have dozens of pairs of wires overhead, and allowed the use of just a few wires. But these few wires still had to be strung somewhere. Underground installation (shielded cable) is seldom a cost-effective solution. With the growth of urban centers in the United States in the 1800s and 1900s, power companies have increasingly found themselves scrambling to find real estate for overhead power lines, particularly for high-voltage transmission lines. Often the best solution has been to lease a portion of the railroads right-of-way. It seems like a natural pairing. The railroads were designed to carry large quantities of goods and raw materials into and out of the cities. They spanned the vast distances between major cities and across rural areas. The growing ac power grid in North America was designed to carry large amounts of electrical power into the cities and across the vast distances that often lay between the generating stations and the cities. Both systems needed a right-of-way network that was hundreds of miles long, and often only a few meters wide. In the increasingly competitive transportation and electric utility markets, the economic efficiency made possible by joint use of railroad corridors has often been an economic necessity. Aware of the potential for compatibility problems, the railroads have usually asked the electric companies to pay for any needed changes or upgrades to their signal system required as a result of interference arising from the power lines. With the increasing complexity, cost, and tort liability associated with railroad signaling systems, this portion of the agreements between railroads and electric utilities has sometimes been a source of considerable friction between otherwise amicable business partners. Of course, railroads can only require such agreements if the power lines are on railroad property. But even if the power line right-of-way is not on railroad property, and is merely adjacent to or near a railroad right-of-way, the railroad can still experience similar levels of interference. In that case, various authorities might assign responsibility. In the United States, state public utility commissions often address such issues, and in some cases, the courts wind up resolving disputes. With thousands of miles of railroads and power lines now in place, the problem of electromagnetic compatibility (EMC) between power lines and railroad signals is one that cannot easily be avoided by the relocation of either. Ensuing problems will have to be solved with much of the equipment remaining in-place. But just as the elements of the compatibility problem between ac power lines and railroad signaling systems come from many sources, so must the solutions. Unfortunately, even though there is a wealth of expertise available in the areas of electric power transmission/distribution, and railroad signaling, there are very few people with experience in solving compatibility problems involving both. To date, there has been only sporadic cooperation and information exchange among the electric utilities, railroads, and railroad signal equipment manufacturers.

1-3

Introduction and Background

In order to facilitate the process of cooperation among the railroads, electric utilities, and railroad signal equipment manufacturers, and to work towards developing effective and standardized solutions to these compatibility problems, the Electric Power Research Institute (EPRI), a nonprofit research institute in the United States founded by a consortium of electric power utilities, has undertaken the task of producing this Power System and Railroad Electromagnetic Compatibility Handbook. For this effort, EPRI has enlisted the assistance of the Association of American Railroads (AAR), the American Railway Engineering and Maintenance-of-Way Association (AREMA), several individual railroads and electric power companies, the railroad signaling equipment manufacturers, as well as leading railroad signal and electromagnetic compatibility experts. While the primary focus is always on safety - ultimately, the motivation for cooperation among all of these parties is economic. The railroads benefit financially by leasing a portion of their property for use by overhead transmission and/or distribution lines. The electric companies benefit financially by getting affordable access to additional routes for their transmission and/or distribution lines. And the signal equipment manufacturers and consultants benefit financially by designing, producing, and selling EMC-compatible equipment that allows for the safe and efficient movement of trains. Shared utility corridors are an economic necessity. It is too expensive and time consuming to develop separate rights-of-way. But these joint use corridors have led to electromagnetic compatibility issues at certain times and in certain places. The most economic solutions to these EMC problems, which have arisen from the sharing of material resources, will come from the sharing of information resources. Good communication, effective investigation techniques, and effective mitigation will benefit the public and all industries involved. This handbook contains the pooled information resources of many corporations, organizations, and individuals, to whom the authors are deeply grateful. It is hoped that this handbook will provide a basis for discussion, education, and problem resolution throughout the electric power, railroad, and signal manufacturing industries.

What is AC Power Interference?


AC power interference in railroad systems is an electromagnetic compatibility (EMC) problem. For an EMC problem to exist three things must be present. There must be a source of the energy (e.g. the power line). There must be a receptor (i.e. the railroad). And, there must be a path. This source-path-receptor model of EMC is an important element in understanding railroad ac interference. The scope of this Power System and Railroad EMC Handbook is limited to situations where the railroad is the receptor circuit (sometimes called the victim circuit). International Electrotechnical Commission (IEC) standards refer to the power system and railroads as sources, and telecommunications systems as victims. These same standards refer to railroads as sources, instead of receptors, because most European railroads use electric traction1. Where diesel electric
1

Electric traction is one of two common ways to propel a train. The other is diesel electric locomotion. All modern trains use electric motors to drive the wheels. A diesel electric locomotive has a diesel engine that drives an electrical generator. The electricity from the generator then drives the electric motors to turn the wheels.

1-4

Introduction and Background

locomotives power the railroad the situation is completely different. A diesel electric railroad is usually a receptor not a source. This handbook will provide processes and tools for identifying, characterizing, and mitigating electric power interference on railroads. In this EMC Handbook, all aspects of ac power interference are addressed. These include effects on equipment operation, damage to equipment, and personnel safety (direct electrical effects on the body).

Prior Programs
Programs to address ac interference challenges on railroads have been undertaken in the past. Most of these programs concentrated on the issues of ac induction. While a study of these prior works is important for a complete knowledge of the field, changes in technology have had a significant effect on the levels of ac interference that can be tolerated. Therefore, the results of prior works must be interpreted carefully. Several examples of these programs follow in chronological order. In 1917, the Joint Committee on Inductive Interference to the Railroad Commission of the State of California published their Final Report in an attempt to address these issues [1]. While on the face of it this report seemed reasonable, a closer look revealed some unsettling issues. In 1922, Mr. John Leisenring of the Illinois Traction System presented a paper to the Joint Convention of the Illinois Gas Association, the Illinois State Electric Association, and the Illinois Electric Railways Association that summed up the utility attitude of the California report and the State regulations that followed. In the paper, titled Inductive Interference [2], Leisenring stated: The much talked of California Rules, prepared but a few years ago, which at the time were heralded as the last word in the matter of joint corrective measures, prepared by a committee of both power and signal men, proved to be inadequate and one sided in that they placed no responsibility on the signal companies that can be definitely applied, while the impossible solution of wide separation is the real basis of the report. Another advocate of somewhat arbitrary rules for the handling of this problem has recently arisen in the form of the Telegraph and Telephone Section of the American Railroad Association. This association, last year, forwarded to every state commission in the Union a copy of the California Rules with the, at least implied, suggestion that they be adopted in each state. I mention this simply to show the need for an immediate and complete awakening on the part of the power interests as to the seriousness of the situation, and urge the hearty support of the National Committee of the National Electric Light Association as well as any State or Regional Committees working on this Problem. Then in 1936 something wonderful happened. The Report of the Joint General Committee of AAR and EEI on The Inductive Coordination of Electrical Supply and Communication Systems, October 7, 1936 was published [3]. This report said things like:
Electric traction uses electricity from another source, usually delivered through an overhead catenary wire or an electrified third rail, to drive the electric motors that turn the wheels. Because electric traction requires the railroads to have their own electric power distribution systems, railroads using electric traction can be sources of ac power interference.

1-5

Introduction and Background

All railroad and electric power systems should be designed, installed and maintained to minimize interference problems. When normal practices are insufficient, the Railroad and Electric Power Company involved should cooperate to solve the problems. This document was last revised in September 1977, and is now known as Principles and Practices for Inductive Coordination of Electric Supply and Railroad Communication/Signal Systems [4]. While it is not a Standard, this twenty-nine year old document, commonly called the Bluebook, was once the closest thing to an industry accepted guide that we had in North America. Although the American Railway Engineering and Maintenance-of-Way Association (AREMA) was working on a revision of the Bluebook, it has been replaced by this handbook. Since 1977, other work was pursued to advance the state of the art of railroad/electric power inductive coordination. Joint projects between the Association of American Railroads (AAR) and the Electric Power Research Institute (EPRI) produced a series of reports and tools: October 1983 Mutual Design of Overhead Transmission Lines and Railroad Communication and Signal Systems, Volume 1: Engineering Analysis and Volume 2: Appendixes, EPRI Report EL-3301, Research Project 1902-1. July 1985 Utility Corridor Design: Transmission Lines, Railroads, and Pipelines, Volume 1: Engineering Analysis and Site Study and Volume 2: Users Manual for Computer Program CORRIDOR, EPRI Report EL-4147, Research Project 1902-2. Companion project to 1902-1 Included development of CORRIDOR software to predict magnetic and electric induction into railroads and pipelines Included creation of the Track Circuit Simulator to permit testing of any interference condition on actual equipment installed on simulated track

This work, as with most of the prior work, concentrated on induction. However, most of the problems railroads have with newer technologies of signal equipment involve interference that is not coupled through induction. Many cases of ac interference are caused by conduction through the earth or other paths. This has lead to a growing number of investigations that are ineffective because they are looking for the wrong kind of problem. It is necessary to broaden the study of ac interference from one of identifying the sources of induction, to an Electromagnetic Compatibility (EMC) study in which all sources and paths are identified. In this way problems can be accurately characterized and effective mitigation implemented in the least amount of time and at the lowest possible cost.

Terminology
Finding unambiguous terminology for this handbook has been a challenge. The audience for this material includes railroad personnel, power company personnel, EMC people, and equipment manufacturers. It includes electrical engineers, mechanical engineers, technicians, signal maintainers, government regulators, and managers. Because of this, every effort will be made to use standard terminology. Where existing terminology standards are insufficient, we define and use our own terms. 1-6

Introduction and Background

The first example of our chosen terminology is ac power interference: When we say ac power interference, we are talking about energy that is undesired. This energy may or may not cause problems to the railroad or its personnel. But the railroad doesnt need this energy (at least not where it is), and if there is too much of it, problems can result. That undesired energy is interference2. We are talking about alternating current (ac). Most systems that use electrical energy to make things go use ac power.3 When we use the word power we mean the electrical energy used to make things go. This is usually 60Hz or 50Hz ac. We use it here as electrical energy at these fundamental frequencies and lower-order harmonics (e.g. multiples of the 60Hz fundamental, such as: 120, 180, 240, 300, 360, 420Hz, etc.).

AC power interference can come from many sources within the power company systems or from sources outside the power company system. Either way we need to be able to identify, characterize, and mitigate ac power interference. So, ac power interference is undesired electrical energy regardless of its source, coupling path, harmonic number, or whether problems result. We do not call it induction. This term is frequently used when ac power interference is present and both the source and path of the energy are unknown. In fact induction is a path actually two different possible paths the energy might take to get to the railroad systems. As will be explained later, it is important not to make the mistake of assuming the path is magnetic induction. Several hundred thousand dollars could be spent mitigating magnetic induction, only to discover that the energy used a different path and that the mitigation is completely ineffective.

Organization
Reference books are usually either written as a textbook that you must read at length to derive the desired knowledge, or as a reference that you can go to for specific things, but cant use to get in depth explanations. This handbook is both - not because we have finally found true enlightenment but simply because we decided to write both books at the same time and bind them together. Many of the chapters of this handbook are written with a quick-start version and a detailed version. For a complete understanding of the topic read the quick-start version and then read the detailed version. If time does not permit this level of study, the quick-start version is designed to provide sufficient understanding to effectively initiate work in the subject area.

In some fields of study the word interference is only used when problems result. They use the word noise to describe undesired energy that does not necessarily cause problems. For reasons described elsewhere, we dont use the term noise here, and interference does not necessarily cause problems. This handbook will also include what we know about the effects of direct current (dc) electric transmission systems, but we use ac as in ac power interference because people understand it, and for some reason people dont understand electric power interference.

1-7

Introduction and Background

References
1. Final Report of the Joint Committee on Inductive Interference to the Railroad Commission of the State of California, California State Printing Office, Sacramento, 1918. 2. Inductive Interference, John Leisenring, Illinois Traction System, Presented before the joint convention of the Illinois Gas Association, the Illinois State Electric Association, and the Illinois Electric Railways Association, Chicago, March 1922. 3. Report of the Joint General Committee of AAR and EEI on The Inductive Coordination of Electrical Supply and Communication Systems, October 7, 1936. 4. Principles and Practices for Inductive Coordination of Electric Supply and Railroad Communication/Signal Systems, A Report of the Joint Committee of the AAR and EEI on Inductive Coordination, September 1977.

1-8

2
FUNDAMENTALS OF ELECTROMAGNETIC COMPATIBILITY (EMC)
This chapter provides an overview of electromagnetic compatibility (EMC) and how electromagnetic interference (EMI) from power systems can affect railroads. It includes interference mechanisms in railway and power situations, common causes of problems, personnel safety considerations, and applicable national and international standards. Jerry Ramie is a 27 year veteran of the Regulatory Compliance, EMC and Microwave instrumentation industries. He is the President and founder of ARC Technical Resources, Inc. which provides training, equipment, systems and services for the Electromagnetic Compatibility (EMC) profession since 1989. He is a Senior Member of the IEEE - EMC Society, the dB Society, the Power Engineering Society P1775 Committee on Broadband over Power line (BPL), the EMC Society's Ad Hoc Committee on BPL and is a NARTE-certified EMC technician. He can be reached at http://www.arctechnical.com Dr. James Stewart is an independent consultant based in Scotia, New York. He has over 30 years of experience in power systems and transmission lines, having worked for Niagara Mohawk Power Corporation and Power Technologies, Inc. He has been involved in experimental and analytical research of compact overhead transmission lines, and is a co-author of the EPRI Transmission Line Reference Book: 115-138 kV Compact Line Design and subsequent EPRI reports on line compaction and DOE reports on high phase order power transmission. This compact line research was subsequently applied to voltage upgrading of existing transmission lines. Dr. Stewart participated in several of these upgrading studies. He was elected as a Fellow of the Institute of Electrical and Electronics Engineers in 1987 for advances in transmission line theory and its reduction to practice through prototype demonstration. He is presently Chairman of the Transmission and Distribution Committee of the IEEE Power Engineering Society.

2-1

Fundamentals of Electromagnetic Compatibility (EMC)

Introduction
Safety and reliability are key concerns for the railroad and power industries. Railroads and power providers both carry massive loads that demand safe, reliable operation. The revenue stream for both industries depends on safe operation and high reliability, and each is becoming more dependent on semiconductor and communication technologies that are less robust than the simpler electromechanical systems of the past. The modernization of railway and power control equipment will increase the incidence of electromagnetic interference (EMI) between these new devices and the existing infrastructure, and/or the increasing numbers of wireless devices used by the public. These trends must be carefully managed to assure the continued safe, reliable operation of railways and power systems. This chapter of the handbook is intended to provide railroad and power personnel with a basic understanding of the following to ensure electromagnetic compatibility (EMC) between railroad and power systems: Electromagnetic interference Interference mechanisms in railway and power situations Common causes of problems Applicable national and international Standards

Quick-Start Version
Introduction to EMC Whenever power lines are in close proximity to railroads, it is necessary to determine whether mutual interference effects are possible. Electromagnetic compatibility (EMC) helps ensure the ability of one system (e.g. railroad signals) to operate in the presence of effects caused by a nearby system (e.g. electric power lines). Such effects include voltages and currents in the railroad facilities caused by unintended coupling to voltages and currents existing on the power system. For electromagnetic compatibility issues to arise, three distinct entities are required: the emitter of interference, the receptor of interference, and the coupling path between them. The coupling path for possibly interfering voltages and currents entering railroad facilities from the nearby power system can involve three distinct but related physical mechanisms: conduction, electric field coupling, and magnetic field coupling. Characteristics of each of these are indicated in the list below. If the coupled voltages are greater than the railroad equipment immunity, improper equipment operation is possible. Voltages and currents are coupled into the rails of railroad tracks in two different modes: common mode and differential mode. Common mode involves equal voltages being present on the two track rails, and differential mode involves different voltages on each rail, resulting in a 2-2

Fundamentals of Electromagnetic Compatibility (EMC)

voltage difference between the rails. These are described in the second table below. Differential mode voltages are generally the more significant sources of interference to railroad signals. Common mode voltages can be converted to differential mode voltages by electrical and physical imbalances between the two rails. Various technical standards and emissions and immunity test procedures have been developed in different parts of the world to assess and deal with EMC concerns. American and European outlooks differ. Legal, regulatory and social issues also arise. There are methods and procedures that can be applied to both the railroad and electric power systems to deal with and mitigate interference concerns wherever they occur.
Table 2-1 Coupling Mechanisms Between Power Lines and Railroads Conductive coupling Conductive path between power line and railroad through the earth Current paths between power system grounded neutrals or shield wires, and railroad facilities grounds Of concern both during normal power system operation and during faults Effects may change with soil moisture content May include effects due to harmonic currents in the power system Electric field coupling Driven by power line voltage Generally results in high open circuit voltage (kV), but only milliamperes of current Receptors are generally pole-top communication wires or long trains parked parallel to power lines Can often be mitigated by grounding methods Tends to be dominant induction path in high impedance pole mounted communication circuits Magnetic field coupling Driven by power line current Generally results in low open circuit voltage (volts), but may be amperes of current Of concern both during normal power system operation and during faults Tends to be dominant induction path in low impedance track signal circuits Concern for possible misoperation of signals, equipment damage and personnel safety May include effects due to harmonic currents in the power system (usually a conduction phenomena)

2-3

Fundamentals of Electromagnetic Compatibility (EMC)

Table 2-2 Common Mode and Differential Mode Coupling Common mode coupling Identical voltage induced per unit length on both rails May be of concern for equipment damage May be of concern for personnel safety Steady-state ac induced in track should not exceed 50 V rms rail-to-ground or across an insulated rail joint Differential mode coupling Voltage difference between rails Of concern for abnormal operation of railroad systems if greater than equipment immunity May be developed from common mode voltage because of unbalanced impedances to ground in track circuit Track circuit balance should be maintained at the best possible level (<10% unbalance, lower if practical) Acceptable levels of differential mode interference are dependent on the immunity of the installed signal equipment. (AREMA Manual Parts recommend 5V and 10V of ac immunity, but many systems cannot operate at these levels of interference.) Grade crossing equipment is particularly susceptible.

Detailed Version
Introduction to EMC Electromagnetic compatibility (EMC) relates to the ability of an electrical device or system to function satisfactorily in the electromagnetic interference (EMI) environment in which it is situated, while at the same time not introducing intolerable electromagnetic disturbances to anything else in its environment. In other words, EMC is the art of controlling EMI and its effects, both from the standpoint of the source and the receptor. Electromagnetic interference can result in unintended or impaired operation of an electrical or electronic system or device whenever interfering voltages or currents exceed the acceptable values. What constitutes an acceptable value is specific to the application under consideration and must take into account the entire setting of the problem. For example: a radio receiver that is designed to receive one frequency in the presence of strong signals on other nearby frequencies has a particular set of design requirements. In this case, the requirements apply to the receiver (the receptor devise), because the interference cannot be eliminated (They are intentional and legitimate transmissions.) On the other hand, a consideration of corona noise from overhead electric power transmission lines is a question of radio signal strength at the line location and the signal-tonoise ratio to be maintained at the edge of the right-of-way. In this case, the requirements

2-4

Fundamentals of Electromagnetic Compatibility (EMC)

apply to the power line (the source), because the only practical way to maintain the required signal-to-noise level is to limit the strength of the interference from the power line. Electromagnetic interference analyses involve not only an engineering analysis of the physical interference mechanism, but also the development and application of criteria for evaluation of the interference potential. If radio noise is an annoyance issue to the public, one set of statistics can be applied at the receiver location. If the issue is safety, such as a malfunctioning grade crossing signal, much more stringent criteria must be used. If the system is not affected by interference, it is said to be immune.

Interference involves three entities: the source (capable of generating electromagnetic energy sometimes called the emitter), the receptor (a.k.a. the victim, which can be harmfully affected by such energy), and the coupling path that allows some amount of interfering energy to be coupled into the receptor from the source. This could be shown as the source-path-receptor model for interference in Figure 2-1.

Figure 2-1 The Source-Path-Receptor Interference Model

Electromagnetic compatibility exists when the source and the receptor both perform their intended function, with the source and path delivering less interfering signal than the receptor is designed to accommodate. Compatibility is thus the state of having sufficiently little interference generated by the source, more than enough immunity to such interference inherent in the design of the receptor, sufficient attenuation in the path, or a combination of these things. Controlling interference is important to assure reliable operations of both rail and power systems. This Handbook strives to present concise, usable information on how to identify interference problems and indicate solutions for them that work.

2-5

Fundamentals of Electromagnetic Compatibility (EMC)

Interference Paths This section discusses the paths for interfering voltages and currents to enter railroad facilities. A subsequent section deals with the principles of how these coupled voltages and currents affect the railroad equipment. There are three general categories of electromagnetic coupling paths between systems: conduction, induction, and radiation. Conduction involves conducting paths, whether intentional through wires, or unintentional such as through the earth or lossy insulation. Induction results from electric and magnetic fields coupling objects that are in close physical proximity to each other. Radiation occurs for more widely spaced objects and involves radio frequency propagation through free space.

Conduction involves a conducting path such as the earth, buried metallic pipes, or wires connecting the source with the receptor. There was a serious issue with conducted earth currents in the early days of the electric street railroad industry. The cars were supplied with DC from an overhead wire with the rails providing the return path to the powerhouse. Because the ballast also provides a conductive path between rails and earth, some of this return current passed through the earth into nearby piping, causing corrosion of the pipes. Methods were developed to mitigate this problem, although there are still concerns today with DC currents in the ground causing corrosion. Power system grounding performs several functions both during normal system operation and during faults. Power system grounds include substation ground mats, distribution system neutrals, transformers and service entrances, and transmission line structure grounds. A multi grounded distribution feeder has grounds on the neutral wire at each transformer, and usually at regular intervals. Transmission line structures are well grounded because the shield wires require frequent grounding to perform their purpose of lightning protection. For a description of the power system itself and a more detailed discussion of power system grounding practices, see the introduction to the power system in Chapter 3. Many conductive connections exist between the power system and the ground, where current enters the earth and can find its way into railroad facilities. Railroads ground (or earth) their signal houses, cases, and bungalows. The ballast also provides a conductive path from the earth to the rail circuit. These provide paths for earth currents from the power system to enter railroad facilities. When power and railroad facilities share rights-of-way, there can be numerous grounds in close proximity, providing a number of possible conducting paths for current between systems. It is important to consider conductive interference both during normal operation of the power system and during faults when very large currents can enter the earth. It is essential to preserve proper operation of railroad facilities during normal power system operation. It is also essential to prevent damage to equipment or injury to personnel as a result of elevated voltages during power system faults. The physical process for conductive coupling between power systems and railroads is illustrated in circuit form in Figure 2-2 and pictorially in Figure 2-3. Resistance in the earth provides a 2-6

Fundamentals of Electromagnetic Compatibility (EMC)

conducting path between the grounds on the power line and incidental connections to ground along the rail. Current flowing through the earth divides between the parallel paths of earth and rail depending on their relative impedances. The relative impedances of the circuit determine the fraction of the current that flows in the earth. Some of those impedances are the resistances connecting the rail to the earth. If the resistance between rail and earth is zero, the division of current between earth and rail is determined by the self and mutual impedances relating earth and rail. A complete analysis includes the matrix of terms for the ground connections themselves.

Figure 2-2 Conductive Coupling Between Power and Railroad Systems

Figure 2-3 Earth Conduction

2-7

Fundamentals of Electromagnetic Compatibility (EMC)

Another approach to considering conductive coupling is through the concept of ground potential rise, shown pictorially in Figure 2-4. Because the earth is an imperfectly conductive body, there is a resistive voltage drop in the earth from one location to another whenever current flows in the earth. For example, a voltage can be measured on a ground rod on a railroad or power system with respect to a ground rod driven in the earth far away (called remote earth). Care must be taken in doing such a measurement because of the possibility of magnetic field induction in the measuring leads introducing an additional voltage into the measurement. However, the concept of ground potential rise is important. A person standing on the earth touching a metallic object may be subjected to a touch potential due to the voltage difference between the earth and the object. Likewise, especially during faults, there is a voltage difference between points on the earth. The difference in voltage between a persons two feet is called step potential. Grounding practices for personnel safety must take these touch and step potentials into account.

Figure 2-4 Ground Potential Rise

Induction is the near-field coupling between closely spaced objects. For instance, inside a transformer magnetic induction is the coupling path for the magnetic field from the primary windings to the secondary windings. What constitutes closely spaced is relative to the frequency and wavelength of the voltage or current. The wavelength () = c / f, where c is 8 the speed of light (3 x 10 m/s) and f is the frequency in Hertz (Hz). For 60 Hz, = 5,000 km, therefore railroad and power system physical dimensions are a small fraction of a wavelength. At these small fractions of a wavelength, the electric and magnetic fields can be considered

2-8

Fundamentals of Electromagnetic Compatibility (EMC)

quasistatic, that is, for many purposes the fields behave as if they were static fields even though they are actually time-varying. In the quasistatic region: Electric field is a function of the source voltage, and Magnetic field is independently a function of the source current.

Thus, electric and magnetic field coupling can be considered separately at power frequencies. The question of electric and magnetic field coupling is fundamental to EMC between power systems and railroads. In this respect it is similar to the broader field of EMC, including compatibility concerns in printed circuit boards. This is because, in electromagnetics, size is measured in wavelengths (), not how big something is in meters or feet. An ideal dipole antenna is about 0.5 long. If the receptor is much less than 1/6 away from the source, it is in the near field. A useful comparison is the dimensions of a printed circuit board compared with the much larger geometry of a joint railroad power right-of-way. A rough equivalence could be viewed as shown in Figure 2-5.

Figure 2-5 Wavelength Scaling when Measured in Wavelengths (), Printed Circuit Boards and Joint Railroad Power Corridors Present the Same Geometry

On a circuit board the length of a source conductor (LS) could easily be 0.01, the length of a receptor trace (LR) 0.001, and the spacing between layers (LPath) could be 0.00001 (f=15MHz, =20m, LS=20cm, LR=2cm, LPath=0.02cm). When measured in wavelengths, a power line operating at 60Hz that is 50km long and 50m away from a track block that is 5km long has the same geometry. In this view, all the theory and practices used to design high-speed circuit boards for minimum emissions and maximum immunity are roughly equivalent to designing rail and power systems for minimum signal degradation and maximum immunity to interference. At

2-9

Fundamentals of Electromagnetic Compatibility (EMC)

either scale, conductors are run in parallel with one another over some distance and are separated by insulators suspended over ground planes.1 Electric field induction is capacitive coupling as shown in Figure 2-6 and is the primary cause of induced voltages on objects the size of motor vehicles, people, sheds, long trains, or pole mounted communication wires. Figure 2-6 illustrates a capacitive voltage divider connecting the line voltage with the object on which voltage is induced. Electric field induction can be represented as a Thevenin equivalent circuit consisting of a voltage source (VOC), an equivalent source impedance (ZS), and the resistance of the object (possibly a person) contacting the parallel conductor (RP), Figure 2-7. The voltage source represents the open circuit voltage induced on the object if it is perfectly insulated from ground. The source impedance determines the short circuit current available. The resulting circuit for electric field induction is a high voltage (on the order of kilovolts), and a source impedance that is the equivalent reactance of the capacitance network (on the order of megohms). Because of the small capacitance between line and object, the capacitive reactance is large. Thus, it is possible for electric field induction to use a Norton equivalent circuit, which consists of a current source (ISC) in parallel with the source impedance (CV and RV in parallel). In this case, it is a very good approximation to use only the current source, and to neglect the source equivalent impedance as shown in Figure 2-8. A pictorial representation is given in Figure 2-9.

Figure 2-6 Capacitive Coupling to an Object Close to an Overhead Line

Note to EMC specialists: Keep in mind, however, that with the extremely long wavelengths of extremely low frequency (ELF) rail and power systems the induction mechanism is electric and magnetic field coupling. RF radiation is not a meaningful coupling path at practical distances, i.e. at less than hundreds of kilometers.

2-10

Fundamentals of Electromagnetic Compatibility (EMC)

ZS

Figure 2-7 Thevenin Equivalent Circuit for Electric Field Coupling

Figure 2-8 Norton Equivalent Circuit for Electric Field Coupling

Figure 2-9 Electric Field Induction

2-11

Fundamentals of Electromagnetic Compatibility (EMC)

In electric field coupling an open circuit voltage of several kilovolts may coexist with a short circuit current on the order of milliamperes. Any conductance to ground reduces the open circuit voltage. The highest possible voltage occurs when there is infinite resistance to ground and only the capacitive reactance connects the object to ground. Short circuit current is the most reliably measured experimental variable because any leakage resistance, even on the order of megohms, reduces the voltage. In the case of electric field induction, the high voltages present on transmission conductors can transfer charge onto the high-impedance, voltage-driven pole-line circuits still used by the railroad industry in many locations. Likewise, nearby telephone circuits could also be susceptible to interference from this path. It is possible to shield electric field coupling by grounded wires placed above the interference receptors. Alternatively, frequency-selective filters can be installed between the communication circuits and the ground to provide a path for 60 Hz to drain to earth. Magnetic field induction is inductive coupling, and is significant for objects that parallel the power line for a significant distance, such as railroads. Magnetic field induced voltage is expressed in volts per unit length along the parallel conductor, and is called the longitudinal electromotive force (LEF, formerly called longitudinal electric field). LEF is a fundamental concept in inductive coordination problems. The equivalent circuit for magnetic field coupling is given in Figure 2-10 for an object parallel to a power line, insulated from earth, grounded at one end, and contacted at the other by an object grounded through resistance. Voc is the total open circuit induced voltage over the length of parallel conductor; Rs and Xs are the real and reactive components of the source impedance; and Rp is the resistance of the object (possibly a person) contacting the parallel conductor. Numerical values of the parameters in Figure 2-10 usually give a low voltage behind a low impedance. The voltage may be less than 100 volts, but the short circuit current may be on the order of several amperes, or tens of amperes. A pictorial representation of magnetic field induction is given in Figure 2-11.

Figure 2-10 Equivalent Circuit for Magnetic Field Coupling

2-12

Fundamentals of Electromagnetic Compatibility (EMC)

Figure 2-11 Magnetic Field Induction

It is important to consider magnetic field induction under both normal operation of the power system and during power system faults. Much higher voltages are induced during faults both because of the greater magnitude of the fault current compared to normal load current, but also because a greater current may be flowing into the earth making a larger inductive flux loop during faults. It may also be necessary to give special attention to harmonic currents when dealing with magnetic field induction. Many electronic devices, such as power supplies and variable speed motor drives, inject harmonic currents into the power system, some to very high harmonic orders. Because mutual reactance is proportional to frequency, 10 amperes at 180 Hz induces three times the voltage that 10 amperes induces at 60 Hz. Also, harmonic currents of three times the fundamental frequency (e.g. for 60 Hz: harmonic frequencies of 180 Hz, 360 Hz, etc.) return via the neutral or ground path, and for that reason produce greater magnetic field induction. It is not uncommon to find induced voltages with very large percentage harmonic content compared to the source current. Because of their high-frequency content, power system transients, such as those resulting from switching operations, can also induce higher than expected voltages into parallel facilities. For example, capacitor switching operations can result in switching transients and consequent induced voltages. Arcing contacts, whether more-or-less intentional, such as those involved with the sliding contact on catenaries or unintentional due to poor connections, can introduce transients and high induced voltages.

2-13

Fundamentals of Electromagnetic Compatibility (EMC)

A complete description of electric and magnetic field coupling to objects in proximity to electric power lines is given in the EPRI publication Transmission Line Reference Book 345 kV and Above [5]. Magnetic induction causes rail-to-ground voltages to appear on both rails simultaneously2. The two rails may have slightly different voltages induced because of the relative geometry of the rails and the power lines, or because of different leakage impedances to earth. The difference between the two rail voltages may cause interference with the track circuit. Magnetic field coupling to signal or telephone wires can be mitigated by closely paralleling the interference receptor wires with a very low resistance well-grounded conductor, shown pictorially in Figure 2-12. Circulating current induced by the magnetic field in the shield conductor (labeled Ground Wire in Figure 2-12) produces a magnetic field that opposes the magnetic field from the power line. The effect is to reduce the magnetic field at the receptor wire location and thus reduce the induced voltage. Radiation is a far-field effect, as opposed to the near-field effect of electric and magnetic field induction. It is essential to distinguish radiation from electric and magnetic field coupling of power frequency voltages and currents. Power frequency interference does not couple through radiation to railroads at significant levels.

Figure 2-12 Shield Conductor for Magnetic Induction


2

This is common mode induction, which is described later in this chapter.

2-14

Fundamentals of Electromagnetic Compatibility (EMC)

The far field condition (much greater than 1/6th of a wavelength3) is when the magnetic and electric lines of flux are orthogonal to each other, and traveling in a coherent electromagnetic wave to the receiver (like a distant radio stations signal). For meaningful radiation to occur, two things are necessary: the receptor must be in the far field of the source, and both the source and receptor must be effective antennas. For the very low frequencies used in steady-state railroad and ac power work, the antennas of either system are too short to be effective radiators or receptors of EM radiation at 50 or 60 Hz. The wavelength of 60 Hz power is 5,000 km (3,000 miles) in free space, and the pole, track, or power lines are relatively short by comparison. And to be in the far field at 50/60 Hz, the railroad would have to be on a different continent from the power line. Therefore, radiation is not a significant form of coupling between ac power lines and railroads. Differential and Common Mode Signal Types When electrical energy is imposed on a circuit, there are two modes that the energy can take: differential mode and common mode. To understand how interference affects railroad systems it is necessary to understand these two modes, how they are used in electrical railway circuits, and how interference affects function and safety. We are going to explain these modes in terms of signals. By signals we dont mean railroad signals. Here we are referring to the more general definition of signals as time-varying voltages or currents used to communicate information. That is, the electrical energy on the circuit, both desired and undesired. The desired signal is the energy that carries the information that the circuit was meant to deliver. This signal can be anything from the energy powering a doorbell, to a bit inside a computer chip, or even the energy in a railroad track circuit. The signals are the good stuff - the information we want. The interference is the rest of the energy in the circuit, from whatever sources.
3

The wavelength () = c / f, where c is the speed of light (3 x 108 m/s), and f is the frequency in Hertz (Hz) or cycles/second. So, for 60 Hz, = 5,000 km. The boundary region between the near field (quasi-static) region and the far field (radiation) region is centered at k * r = 1, where k is the wave number (k = 2 / ) and r is the distance (m). The near field region (where quasi-static effects dominate) is at distances where k times r is much less than 1 (k * r << 1). Approximating 2 as 6, and solving for r gives r << 1/6 for the near field. So, if your distance is much less than 1/6 , you are in the near field. (A factor of 3 is often used for much less/more than leading to r < 1/18 . This is often rounded to r < 1/20 as the near field.) Conversely, the far field region (where radiation dominates) is r >> 1/6 . A factor of 3 for much greater than gives the far field as r > . Many sources use a factor of 6 resulting in a far field definition of r > . So, with a 60 Hz power line, less than 250 kilometers distance is definitely near field, and greater than 5,000 kilometers is definitely far field. Still want more?! OK. When someone points out that this is true for free space, but what about the wavelength in a conductor such as the wire isnt it shorter? The answer is: yes, but not by much. The wavelength is a function of the material in which the fields exist. For a perfect conductor the fields are entirely outside the wire (in free space). But, wire is not a perfect conductor; so about 5% of the fields exist inside a copper wire. Since 95% of the fields exist in free space, the free space wavelength can be safely used. What about the wavelength in the earth, such as a pipeline? The wavelength in the earth is about one skin depth (DS). Skin depth in a reasonably conductive material is: 1 where: for non-magnetic material ~ 4 x 10-7 DS = for typical earth ~ 10-2 (f)1/2 Therefore, Earth ~ 650 meters

2-15

Fundamentals of Electromagnetic Compatibility (EMC)

Every circuit must have a continuous path for the current to follow, in order to be a complete circuit, as in Figure 2-13. Digital logic sometimes appears to be a single wire, but there is always a ground reference that is the return path. So, from the source end of the circuit we see two conductors heading off into the distance to deliver information (energy) from the source to the load. How the source applies our signal to these two conductors determines whether the signal is differential mode or common mode.

Figure 2-13 Complete Circuit with Source, Path to the Load, Load, and Path from the Load

Differential Mode Signal Differential mode signals are different from each other, that is, out of phase or showing opposite polarity at any given instant in time. For a circuit driven by a battery, the positive terminal is connected to one conductor and the negative terminal is connected to the other conductor, Figure 2-14. This is a differential mode signal. In the circuit that powers the outlet your computer is plugged into, the hot wire and the neutral wire come from different taps on the transformer secondary winding. So, they both carry an ac current, but the currents are 180 degrees out of phase. This is also a differential mode signal.

2-16

Fundamentals of Electromagnetic Compatibility (EMC)

Figure 2-14 Circuit with Differential Mode Signal

Common Mode Signal With common mode signals both conductors receive the same excitation. The same voltage is applied to both. For this to make sense, there needs to be some reference. If you hook both wires of a circuit to one terminal of a battery, but connect nothing to the other terminal, nothing happens. The reference we often use is called remote earth. Remote earth is presumed to be at zero Volts. So, if you take your 24 Volt battery and hook both conductors to the positive terminal, and connect the negative terminal to the earth, then both conductors are energized at +24 Volts, Figure 2-15. This is a common mode signal. As no current flows in this circuit from the common mode signal, it appears to be a very inefficient means for delivering information. If there were leakage resistance to ground along the path, a common mode current would flow in the two conductors. In reality all conductors are grounded, at least through parasitic capacitance. This capacitance to ground makes some complete common mode path possible. Because capacitive reactance decreases with increasing frequency, a capacitance-completed common mode path has lower impedance at higher harmonic frequencies, and is therefore more effective.

2-17

Fundamentals of Electromagnetic Compatibility (EMC)

Figure 2-15 Circuit with Common Mode Signal

Interference Types Now that we know what differential mode and common mode signals are, we need to learn about differential mode and common mode interference, and how they impact electromagnetic compatibility. Common Mode Interference Most circuits that carry information (signals) have their two conductors close together. In many applications they are even twisted together. There is a good reason for this. It means that whatever electromagnetic energy couples into one wire via magnetic or electric induction will be nearly the same as what couples into the other wire (for a well balanced circuit). With the same excitation applied to both wires this induced interference is common mode. Common mode interference does not result in differential current flow in the receptor circuit, Figure 2-16. For this reason it usually does not interfere directly with operation of the receptor circuit. However, common mode interference can cause some problems. If the voltage exceeds the insulating capability of any components, they can fail. Also, if the energy in the receptor circuit exceeds levels that are safe for human exposure, then a dangerous situation could result.

2-18

Fundamentals of Electromagnetic Compatibility (EMC)

Figure 2-16 Circuit with Common Mode Interference and Differential Mode Signal

Differential Mode Interference Differential mode interference is the case where the undesired energy on the two conductors is not the same. It can differ in magnitude, in phase, or in frequency or waveform. Any time that the undesired energy on the two conductors is not exactly the same - this is called differential mode interference. Differential mode interference has the ability to disturb the operation of railroad signal equipment at interference voltage/current levels far below those needed for common mode interference to cause problems. The currents that result in the receptor circuit from differential mode interference can make it difficult for receivers (i.e. the load) to detect their intended signal. Differential mode interference comes about in two ways. First, if the conductors are separated by some distance, then they can be exposed to different fields and have different voltages and currents induced. (This can happen inside a poorly wired equipment enclosure as well as on the track.) However, in the case of railroad track rail spacing, this directly induced differential mode interference is usually not significant. The differences in field exposure and earth conduction exposure are almost always very small for the rail gauge spacing of about five feet. However, differential mode interference should be considered when high-current power lines are closely paralleling the track. The other way differential mode interference gets into circuits is for common mode interference to be converted to differential mode interference by unbalance of the receptor circuit. An extreme (but fairly common) example of this is shown in figure 2-17. Common mode 2-19

Fundamentals of Electromagnetic Compatibility (EMC)

interference of 4 V ac rms exists on both conductors. The differential mode interference is 0 V if there is no connection to ground. Then one conductor is grounded, creating an extreme circuit unbalance. Now with one conductor at 4 V ac rms and the other at 0 V, the differential mode interference is now 4 V ac rms. This simple analysis ignores the source impedance. If the source impedance were zero, there would be a dead short circuit across the voltage source. Current would actually be limited by the impedance of the current loop. However, this still illustrates the extreme case where a common mode voltage is entirely converted to differential mode.

Figure 2-17 Common Mode Voltage Converted to Differential Mode

With varying degrees of receptor circuit unbalance caused by either source impedance differences to the two rails, or differing resistance to ground (Figure 2-18), the differential mode interference can vary anywhere from 0 V to 100% of the common mode interference level. When railroad signal problems occur due to differential mode interference, this interference usually results from a combination of common mode coupling and receptor circuit unbalance, Figure 2-19. Different leakage resistance to ground from the two rails through the ballast is a form of unbalanced impedance that can result in differential mode interference.

2-20

Fundamentals of Electromagnetic Compatibility (EMC)

Figure 2-18 Circuit with Different Resistances to Ground

Figure 2-19 Circuit with Unbalance: Common Mode Interference Converted into Differential Mode Interference

Receptor Circuit Unbalance The diagram in Figure 2-20 shows the relative polarity of both differential mode and common mode voltages in the railroad signaling system. For proper operation of railroad signals, differential mode is desired in the track or pole-line circuits. These track or pole-line circuits require good circuit balance for the signaling voltages to reach their destination without undue 2-21

Fundamentals of Electromagnetic Compatibility (EMC)

distortion. Balance of the line implies that the resistance, capacitance and inductance of each conductor should be the same with respect to ground. If an unbalancing condition exists, that is, if the railroad control or signal lines are not balanced for some reason, interference with the railroad control or signal system will be the likely result.

Figure 2-20 Differential Mode and Common Mode Voltages on Railroad Signals

There are several causes of unbalanced conditions in rail circuits within the railroad control or signaling system itself. Several major unbalanced conditions and their causes are: Open wires in track or pole-line circuit Wires in track or pole-line circuit with abnormally low resistance to ground Abnormally low resistance from one or both rails to ground Failed rail-to-ground surge protective devices (SPDs, a.k.a. lightning arresters) Shorted insulated joints (results in unequal rail lengths)

If one of the lines to the track circuit is grounded, creating an unbalance, the common-mode interference from earth conduction or either induction path still gets impressed on the rails. With one line held to ground potential (0 volts), this common-mode energy now appears as differential across the control or signal input terminals, and interference with the rail system becomes possible.

2-22

Fundamentals of Electromagnetic Compatibility (EMC)

European and North American Models The railroad industry employs two types of locomotion: 1. Self-powered (usually diesel-electric), and 2. Electrified (third or fourth rail, overhead wire or catenary). Electrified locomotives are typical in the European Union and many other countries while diesel electric locomotion is more widely used in the Americas. European Model The European model for railroad interference could be viewed as shown in Figure 2-21.

Figure 2-21 European Interference Model

In this view, the source of EMI is the electrified rail network or ac power equipment, and the receptors are the telecommunication equipment located nearby. The three major types of coupling (magnetic and electric induction and earth conduction) are listed. In Europe the EMC Directive of the European Union (89/336EEC) mandates that emissions shall be kept below the legal levels specified, and that the equipment shall demonstrate an intrinsic (designed in) resistance to outside interference (as the relevant testing standards demonstrate). This logic attempts to ensure electromagnetic compatibility by pre-deployment testing of the equipment or subsystems to recognized international standards for both emissions and immunity. Known railroad sources of interference would include transient magnetic fields and conducted switching transients from the switching operations in the rail substations. Induced and conducted harmonics are produced by the non-linear rectification used to produce the dc and traction ac 2-23

Fundamentals of Electromagnetic Compatibility (EMC)

train power at these installations. As with all power substations, corona, noise from dirty insulators, and gap discharges between metallic parts all contribute to the local electromagnetic environment and provide a coupling path to the outside power, wireless and telephone systems. [6] North American Model In the U.S., and other diesel-electric rail areas, the North American model4 for interference mechanisms is more relevant, as shown in Figure 2-22. The source of interference energy is shown as the distribution or transmission equipment used in the electric power system, and the receptors are the track or pole-line used by the railroad, or telecommunication equipment. The same major coupling paths of magnetic and electric induction and earth conduction are also listed.

Figure 2-22 North American Interference Model

The source-path-receptor model can be viewed as European or North American in nature, depending on what is the source and what is the receptor. In either model, the dominant coupling paths are magnetic induction, electric induction, and earth conduction. In either model, a common-mode voltage can result on the railroad control or signal lines from any of these three paths. This common-mode interference voltage is usually ignored by the railroad equipment, which has high common-mode voltage rejection. However, if an unbalancing condition is present it can force this common-mode interference to be impressed as differential-mode interference onto the leads of the control or signaling equipment, often causing abnormal operation.
4

It is important to note that the European Model is applicable wherever electric traction is used, and that the North American Model might be appropriate wherever diesel electric locomotion is used. The name European Model is used because European standards presume electric traction. The name North American Model is used because most trains in the Americas do not use electric traction.

2-24

Fundamentals of Electromagnetic Compatibility (EMC)

Personnel Safety All employees of electric power companies may be exposed to high levels of extremely low frequency (ELF) power-frequency electric and magnetic fields. The United States sets limits on this field exposure through OSHA, and the European Union sets exposure limits through the IEC, CENELEC and other standards organizations. The issue of personnel safety from direct exposure to magnetic fields, electric fields, and electrical potentials on exposed conductors is addressed in Chapter 7. Emissions & Immunity Testing The coupling mechanisms most prominent in modern shared rail and power environments are: Magnetic Induction Electric Induction Earth Conduction

Weve mentioned that interference consists of emissions coming from sources and ending up as problems in receptors. Traditionally, the coupling paths for this interference were either conducted (through wires or earth) or induced (or radiated) through free space. This can be visualized as a simple 2x2 matrix of tests needed to determine if a product or system was electromagnetically compatible. A simple way of labeling these kinds of tests is shown below as the matrix of Table 2-3.
Table 2-3 Essential EMC Tests Emissions Conducted Induced (or Radiated) Conducted Emissions Radiated Emissions5 Immunity Conducted Immunity Radiated Immunity

Conducted emissions, radiated emissions, conducted immunity, and radiated immunity are the four essential types of tests. These tests are performed to ensure that source emissions do not exceed receptor immunity. Any piece of equipment used in the shared environment of rail and power systems could be both a source and a receptor of emissions. It is important in any complex distributed system like the shared environment of railroads and power systems that electromagnetic compatibility is assured. Causes of interference should be addressed using the most expeditious and cost-effective method, regardless of where the problem may reside or manifest itself. Electromagnetic interference (EMI) is a fact of modern life, with many new sources of interference and new receptors being introduced all the time. Sources
5

The EMC community uses radiated to include both induced and radiated energy. We use their term here because EMC standards use this term. Remember that for power-frequency interference on railroads we are primarily concerned with induced energy, both electric and magnetic.

2-25

Fundamentals of Electromagnetic Compatibility (EMC)

need to be controlled and monitored for changes in their emissions over time. Immunity to interference needs to be assured so that the systems may function reliably in the environment for which they were intended. Each industry needs to understand their sources and their receptors. Each industry should take appropriate action to monitor and address them, thus assuring the electromagnetic compatibility needed in critical systems like railroads and power systems [7]. Types of Emissions Tests Radiated emissions tests set limits on how much RF energy a product can radiate through space. The tests are typically run with a fixed receiving antenna located 10 meters from the product being tested. The product sits on a turntable and is rotated to obtain the maximum emissions. Larger products are tested on all sides, and the field strengths compared to appropriate limits. The site for radiated emissions testing can be an open area test site (OATS) or a 10-Meter anechoic chamber, both installations being without radio signals or reflections [8]. Conducted emissions tests measure how much energy the product can leak back into (and possibly re-radiate from) its power cord. The product is connected to a line-sampling device that senses the radio noise that is coming from the product and delivers it to a measurement receiver where its compared to the allowed limits. Also, the shape of the current demanded by the product under test can be seen as a power-line emission. If the current waveform is non-sinusoidal, such as a switched-mode power supply demands, it will be rich in harmonic content. These harmonic emissions (currents) will distort the main electric lines and affect distribution transformers by adding low-order odd harmonic power to their neutral lines, sometimes causing overheating and damage [9]. How often and how deeply the product draws current from the main power lines can affect other devices nearby on single-phase lines. The varying amplitude of current being demanded by a product can cause differing brightness from incandescent lamps connected to the same lines. This flickering of the lighting is considered a flicker emission and is regulated in Europe as a form of power-line related emissions under the EMC Directive [10]. Types of Immunity Tests For higher reliability prediction of the potential for interference, immunity testing can be used to quantify the products resistance to outside interference, and assumes several forms that represent known threats to electrical & electronic equipment. Continuous RF energy radiating onto the chassis of a system can occur in urban areas with strong local transmitters. Such energy or the effects of its modulation could affect high-reliability products, and testing for radiated immunity ensures compatibility in the intended harsh environment [11]. To simulate the effects of radiated RF energy from local transmitters onto the power, I/O, or data lines of high-reliability distributed installations requires conducted immunity testing of those lines. Conducted immunity methods are used for this simulation because they are repeatable, using capacitive or inductive clamp injection techniques. These methods place amplified RF voltages onto the power & I/O lines of the equipment to assess the resistance to induced common-mode interference voltages [12]. 2-26

Fundamentals of Electromagnetic Compatibility (EMC)

Immunity Performance Immunity to any type of interference is quantified by monitoring the equipment under test (EUT or product) during a test for failures of the types shown in Table 2-4 [13].
Table 2-4 Performance (failure) Criteria Type A B C D Criteria No failures, worked normally during test Self-recovering upset, works after test Non self-recovering upset, needs resetting Non recoverable (damaged)

Power-line related tests try to simulate several effects such as power-frequency magnetic fields, impulse magnetic fields and damped oscillatory magnetic fields that are threats in heavy industrial settings like rail or power installations [14, 15, 16] Other line interference tests simulate line dips and interrupts, switching oscillations, harmonic and inter-harmonic immunities, and low frequency conducted disturbances. [17, 18, 19, 20] These power problems are simulated with special transient generators and programmable ac sources while observing the product for the signs of failure such as those in Table 2-4. Other threats are transient in nature like static discharge, arcing switch contacts, or nearby lightning storms. Humans (users and service people) are also a threat to modern MOS-based circuitry because they can transfer charge into the product through touch, in the form of electrostatic discharge (ESD) [21]. Fast rise-time, low-energy conducted events on power and I/O lines can result from switch arcing, and are simulated with electrical fast transient (EFT) testing on power and I/O lines [22]. The electromagnetic effects of lightning strikes can be destructive to wire-connected products within about five miles. To simulate the damage inflicted by these indirect strikes, surge testing is conducted on power, I/O and telecommunications lines [23]. Regulatory Structure There are two world-views on interference, with two regulatory environments that reflect them: the American and European Union structures. In the American view, emitting electromagnetic energy is acceptable as long as any interference thus generated is not harmful. That is, if no harmful interference is generated, the situation is legal. If the interference generated is harmful or interferes with licensed services, the situation may require equipment modifications or mitigation of the interfering condition before operation resumes. In the European view, both emissions and immunity performance of rail or power equipment is to be assured by testing before the equipment is placed into service. A competent body assembles a technical file that includes an EMC management plan. These plans include testing each subsystem to appropriate EMC standards, and then testing the whole vehicle/assembly or power installation for emissions (if practical). 2-27

Fundamentals of Electromagnetic Compatibility (EMC)

Basic EMC Standards The fundamental types of EMC tests were given in Table 2-3, and include conducted emissions, radiated emissions, conducted immunity and radiated immunity. By extending this simple model to the other types of interference widely responsible for equipment malfunctions, a larger matrix can be developed. The tests shown are written for European, military, or U.S. types of equipment, and they reference the type of disturbance being tested for (conducted emissions, for example). These basic standards describing emissions and immunity characteristics displayed by real-world electronic equipment and systems are shown in Table 2-5.
Table 2-5 Basic EMC Standards Type of Disturbance Radiated (radio waves) Conducted (in wires) Power-line related Emissions Military, U.S., Europe Military, U.S., Europe Europe, harmonics/flicker Immunity Military, Europe Military, Europe Military, Europe, magnetic fields, dips & interrupts, oscillatory waves, interharmonic tolerance, common-mode disturbances Europe Europe Europe, & some U.S. testing

Electrostatic Discharge Electrical Fast Transients Lightning (surge)

Standards have been written to quantify these various characteristics and to measure how little a product emits (emissions tests) and to predict how well it will resist several forms of outside interference, or how immune it is (immunity tests). EMC Standards in other Industries Eleven immunity tests, and the four emissions tests make up the basic standards for EMC testing in the European Union (EU), Table 2-5. These standards are also the basis for voluntary EMC standards used by providers of power installations and medical products in the EU and U.S. These basic standards are also the fundamental tests for assessing EMC performance of railway components and subsystems in the European Union. A matrix of these EMC standards requirements might look like Table 2-6.

2-28

Fundamentals of Electromagnetic Compatibility (EMC)

Table 2-6 EMC Requirements Using Basic Standards Power installations (U.S.)

Power installations (EU)

Medical Products (U.S.)

Medical Products (EU)

Railway C&S (U.S.)

Railway C&S (EU)

EMC Requirements using Basic Standards

Emissions

Immunity

D D D D D D Radiated Emissions D D D D D D Conducted Emissions D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D D Harmonics & Flicker Emissions ESD (Electrostatic Discharge) Radiated RF fields EFT (Electrical Fast Transient) Surge (lightning strike) Conducted RF Voltages Power-frequency Magnetic Fields Pulsed Magnetic Fields Oscillatory Magnetic Fields Line dips, interrupts and variations Harmonic & interharmonic tolerance Common-mode line disturbance

All these EMC product standards call out similar basic standards for emissions & immunity to assess conformity, except AREMA (U.S. railways), which only requires RF emissions and surge immunity testing. Also note that the European and American aircraft and automobile industries also use their own EMC testing standards for emissions and immunity that are internationally recognized. The railway rolling stock and control & signaling industry in the U.S. (through AREMA) only imposes EMI radiated emissions testing to FCC Part 15 on rolling stock and radiated & conducted emissions limits on wayside equipment. Radiated emissions and conducted emissions are required on more sophisticated wayside equipment using microprocessors. The limit lines for radiated emissions from microprocessor-based railroad control and signaling equipment in the U.S. are shown in Figure 2-23.

2-29

Fundamentals of Electromagnetic Compatibility (EMC)

Radiated Emission Limits, AREMA vs. FCC at 10 m


AREMA Limit 400 FCC Limit

350

300

250

V/m

200

150

100

50

0 0.01

0.1

1 MHz

10

100

1000

Figure 2-23 FCC Part 15 Radiated Emissions Limits for Digital Devices

The red limit line shows the maximum E-field strength permitted for radiated emissions from ordinary digital devices in the U.S., measured with average detection. The blue line shows the limits established for North American railway digital devices, such as control and signaling equipment with microprocessors. This railroad digital device limit is measured with peak detection, capturing the maximum emission amplitudes. This radiated emissions limit is extended down to 50 kHz to prevent conducted low frequencies from inducing interference into nearby control or signaling circuits. AREMA also specifies the lightning surge protection required in North America for several railway applications, including primary, secondary, and tertiary levels for track, pole-line, and power circuits [24, 25, 26, 27]. Surge is one of several immunity standards in the EUs basic standards series. No other EMC immunity standards are required in the U.S. for control or signaling equipment. Regarding power installations, those in the U.S. have probably not been installed or tested to CIGRE 36.04, since the Standard is new and voluntary (optional). The same could be said of most power installations in Europe, which were also installed before the EMC Directive mandated EN 61000-6-5 emissions & immunity tests. U.S. and EU legal environments are also different in regards to EMC enforcement. 2-30

Fundamentals of Electromagnetic Compatibility (EMC)

Legal Environments In the U.S., the concept of harmful interference takes precedence, stated in the Code of Federal Regulations (47CFR Part 15). The various parts of Title 47 affect intentional and unintentional radiators, as well as incidental radiators such as power equipment. The law requires operation to cease if the product or installation is causing harmful interference to licensed services, such as amateur radio, broadcast, TV, etc. The operators of the interfering device shall remedy the interference condition at their own expense before the resumption of its operation [28]. In the EU, the concept of compliance is upheld. The EMC Directive, 89/336 EEC, states that emissions shall be kept below legal limits and that the products have intrinsic (designed in) immunity to outside interference to enable it to function properly in its intended environment. A manufacturers compliance with this new approach Directive is indicated by testing to all relevant EMC emissions and immunity basic standards and stating such in the Declaration of Conformity that accompanies each product shipped with the required CE Mark. At first glance, it would appear that Europe has legislated away EMC problems between the railway and power systems, but that isnt exactly the case. There are many national systems in place, much of which was installed before there were EMC standards written for them, and old technical barriers still remain. In the U.S., there are some emissions standards in force to avoid the creation of harmful interference, but immunity standards are not universally applied. Each industry, such as computers, telecommunications, medical products, industrial process & control, scientific measurement, power products, aircraft, autos, etc. has had to impose voluntary immunity standards upon itself. Often, U.S. industries have simply adopted the European immunity requirements that they had to meet anyway to sell their products into the EU. This informal system has left some gaps in EMC standards for the railway and power industries in the United States: Power installations are not currently tested to CIGRE 36.04 EMC Standards Electrified railways & their power stations do not have EMC standards in place Railway control & signaling equipment in the U.S. is tested to very few immunity standards

Social Environments Changes in society have caused decreases in the number of new wired telephone lines being installed, thus announcing the rise of broadband and wireless technologies, including fiber optics. The chips that run everything from clocks to wayside equipment and power plants are becoming smaller inside. This shrinking geometry allows the addition of more features and capabilities, and their lower power consumption uses smaller voltage signals. These new devices always feature faster clock speeds and wider-bandwidth I/O ports including video and wireless (microwave) signals. All these trends point to increasing occurrences of EMC problems with and between these devices and the transportation electronic infrastructure.

2-31

Fundamentals of Electromagnetic Compatibility (EMC)

In a transportation setting, these trends come with the new worldwide threat; the uncontrolled passenger. In commercial aircraft, all of us are familiar with the warnings that flight attendants give before take-off to turn off your cell phone and computer, and most of us do so. In railway (and bus) transportation, the passenger expects to use intentional radiators such as cell phones (in several bands, some with 2-way capability) and wireless connections such as Bluetooth or WiFi (IEEE-802.11b) with laptop computers or personal digital assistants. Some PDAs even offer 2-way wireless e-mail and gaming, and all CD players and other personal electronics emit RF energy that could interfere with some aspects of passenger rail operations. Because of the increasing reliance on chips and broadband wireless technologies, all industries, particularly telecommunications, pipelines, railroads and power installations are vulnerable to intentional EMI or sabotage because of their long conductors (acting as antennas). De-commissioned military equipment is available throughout the world at auction prices and some has been re-purposed for bad intent. Gambling machines have been compromised, engine control computers fried, and other instances of high power microwave damage have been documented [29]. Addressing the gaps Are these gaps an American problem? Clearly, market forces have been slow in driving vendors to adopt voluntary EMC standards, as the U.S. medical device industry experience has shown. Problems with electric wheelchairs brought about the adoption of European-style EMC basic standards by the FDA in 1995. Recalls & problems have thus defined the edges of progress against interference conditions, which is a little sloppy and somewhat uncertain. In effect, the doctrine of harmful interference allows the building of a system first and then its evaluation for EMC performance afterwards. (Particularly for the various Immunity threats in Table 2-6.) The European compliance approach has the benefit of orderly implementation of prudent testing standards for all consumer, commercial and light industrial products to ensure compatibility (reliability). Heavy industrial products are now being addressed in the power and railway industries, but much of their existing infrastructure was installed previously. Large, distributed systems like rail & power cannot be tested before they are built, only subsystems testing is possible. The competent body that assembles the technical file for the complete system for the railroad or power provider drafts an EMC management plan, but success is still determined after the completed system is installed. Many manufacturers have complained of a greater market barrier to entry into the EU posed by these basic standards, even those within Europe. They complain about the higher development & unit costs associated with testing to these basic standards and designing their products to pass them. In high-reliability life critical industries, however, these arguments are somewhat flat. The lesson of the past eight years of the EMC Directive in the EU is that orderly transition to emissions and immunity standards testing is better than no transition. Top-down imposition of an EMC Directive chafes many people, but bottom-up implementation of change (improvement) is slow to materialize and often follows problems. Neither system is perfect.

2-32

Fundamentals of Electromagnetic Compatibility (EMC)

Summary We currently have two systems of EMC standards in the world that are in the process of harmonization in many industries. Many new European standards were based on existing American standards, and vice versa. World-designed products are becoming the norm, which can be sold in any part of the world quickly, decreasing their time to market and increasing their world market share. In the EU, rolling stock and wayside equipment is being tested for emissions and immunity before being CE marked, as required. European power providers now acquire equipment tested to essentially the same basic standards suite. In the U.S., railway equipment is tested for FCC radiated & conducted emissions and ANSI surge only, omitting the other threats simulated in the basic standards. Power providers have a CIGRE standard (using the basic standards suite) that was rarely used in their acquisitions, until recently. Neither railways nor power systems are hardened to intentional EMI (IEMI) or sabotage anywhere in the world, other than some military installations. In Europe, further work is underway on the interoperability of the various national railways. In the US, more harmonization with the basic standards would enhance compatibility and reliability in the railway and power equipment industries. AREMA could be the framework for further implementation of the basic standards for railway equipment. Limits or methods may require adjustments, but a recognized International EMC standards structure could then be established. The same idea could apply to the implementation of the basic standards through CIGRE for power equipment sold in the U.S.. These trends are already emerging since railway and power equipment manufacturers have already installed EMC testing facilities in the U.S. These facilities are featured in their sales presentations, citing enhanced reliability benefits. For the existing railway and power installations in Europe or America, improvement in electromagnetic compatibility (EMC) requires each industry to address its sources and its receptors, and to respect those of the other industry. Railways Need To: Identify counterparts in the power industry and maintain communication with them Incorporate EMC considerations in the planning and design of facilities Insist on EMC basic standards testing for new equipment Maintain good electrical balance in rail and pole-line circuits Provide training for staff on interference diagnosis & correction techniques

Power Providers Need To: Identify counterparts in the railroad industry and maintain communication with them Incorporate EMC considerations in the planning and design of facilities Insist on EMC basic standards testing for new equipment Maintain good balance of currents on distribution lines to the extent practical 2-33

Fundamentals of Electromagnetic Compatibility (EMC)

Maintain distribution line neutral conductors in good condition to reduce earth injection Consider mitigation for harmonic currents Sense and diagnose defective capacitor banks Consider mitigation for electrical noise from switching operations Provide training for staff on interference diagnosis & correction techniques

Railway Signal Equipment Manufacturers need to: Incorporate EMC considerations in the planning and design of equipment Comply with voluntary EMC basic standards testing for new equipment Develop new technologies to replace existing grade crossing train detection systems

Systems-level EMC is becoming more important in design and operations of the railroad and power installations. Trends in device physics and control electronics are causing a proliferation of many new EMI sources and receptors, and electromagnetic compatibility will be a growing concern in the future. With the increasing use of semiconductor-based control equipment, the reliability or safety of the railway and power infrastructure could be compromised if EMC is ignored rather than managed. In response to this need, world market forces in all large industries are encouraging the adoption of a single, harmonized set of EMC (and other relevant) standards for world-designed products and systems. The management of systems-level EMC requires a plan and a standardized testing regimen to validate design decisions and the EMC performance of the subsystems. Testing and mitigation of the completed systems is still required to assure electromagnetic compatibility within and between the finished railroad and power systems.

References
5. Transmission Line Reference Book 345 kV and Above: Third Edition, EPRI, Palo Alto, CA: 2005. 1011974. 6. Safety critical EMC on railways, John Whaley & Dave Kearney, Approval Magazine, July/August 2001. 7. Principles & Practices for Inductive Coordination of Electric Supply and Railroad Communication/Signal Systems, Association of American Railroads & Edison Electric Institute, September 1977. 8. IEC/CISPR 16-1, Radio disturbance & immunity measuring apparatus, International Electrotechnical Commission, 1993-08. 9. IEC/EN 61000-3-2/A14, Limits for harmonic current emissions, CENELEC, October 2000. 10. IEC/EN 61000-3-3, Limitation of voltage fluctuations and flicker in low-voltage supply systems, CENELEC, 1995. 11. IEC 61000-4-3, Radiated, radio-frequency, electromagnetic field immunity test, st 1 Ed., International Electrotechnical Commission, 1998-11. 2-34

Fundamentals of Electromagnetic Compatibility (EMC)

12. IEC 61000-4-6, Immunity to conducted disturbances, induced by radio frequency fields 1st Ed., International Electrotechnical Commission, 1996-04. 13. IEC/EN 55024, Information technology equipment Immunity characteristics, Limits and methods of measurement, CENELEC, September 1998.
st 14. IEC 61000-4-8, Power frequency magnetic field immunity test 1 Ed., International Electrotechnical Commission, 2001-03.

15. IEC 61000-4-9, Pulse magnetic field immunity test 1st Ed., International Electrotechnical Commission, 2001-03.
st 16. IEC 61000-4-10, Damped oscillatory magnetic field immunity test 1 Ed., International Electrotechnical Commission, 2001-03.

17. IEC 61000-4-11, Voltage dips, short interruptions and voltage variations immunity tests 1st Ed., International Electrotechnical Commission, 2001-03.
st 18. IEC 61000-4-12, Oscillatory waves immunity test 1 Ed., International Electrotechnical Commission, 1995-05.

19. IEC 61000-4-13, Harmonics and interharmonics including mains signaling at a.c. power port 1st Ed., International Electrotechnical Commission, 2001. 20. IEC 61000-4-16, Test for immunity to conducted, common mode disturbances in the frequency range 0 Hz to 150 kHz 1st Ed., International Electrotechnical Commission, 1998-01. 21. IEC 61000-4-2, Electrostatic discharge immunity test 1st Ed., International Electrotechnical Commission, 1999-05. 22. IEC 61000-4-4, Electrical fast transient/burst immunity test 1st Ed., International Electrotechnical Commission, 1995-01. 23. IEC 61000-4-5, Surge immunity test 1st Ed., International Electrotechnical Commission, 1995-02. 24. Recommended Design Criteria and Function of Primary Surge Protective Devices for Communication and Signal Systems, Part 11.3.1, AREMA C&S Manual, Revised 2003. 25. Recommended Design Criteria and Functional/Operating Guidelines for Secondary Surge Protectors for Electrical Surge Protection of Signal Systems, Part 11.3.2, AREMA C&S Manual, Revised 1996. 26. Recommended Design Criteria for Surge Withstand Capability of Electronic Signal Equipment for Signal Systems, Part 11.3.3, AREMA C&S Manual, Revised 1996. 27. Recommended Design Criteria and Function of Solid State AC Primary Surge Protective Devices (SPDs) for Communication and Signal Systems, Part 11.3.6, AREMA C&S Manual, Revised 2003. 28. Code of Federal Regulations, Title 47, Volume 1, Part 15 Radio Frequency Devices U.S. Govt. Printing Office, revised October 1, 1999 47CFR15.33 pages 664-665. 29. Crossed Signals The wireless threat to our electronic infrastructure, U.S. News & World Report, December 16, 2002 (pages 54-56).

2-35

3
OVERVIEW OF POWER SYSTEMS
This chapter provides an overview of the electric power transmission and distribution systems, with specific concentration on the factors impacting electromagnetic effects on railroads. Dr. James Stewart is an independent consultant based in Scotia, New York. He has over 30 years of experience in power systems and transmission lines, having worked for Niagara Mohawk Power Corporation and Power Technologies, Inc. He has been involved in experimental and analytical research of compact overhead transmission lines, and is a co-author of the EPRI Transmission Line Reference Book: 115-138 kV Compact Line Design and subsequent EPRI reports on line compaction and DOE reports on high phase order power transmission. This compact line research was subsequently applied to voltage upgrading of existing transmission lines. Dr. Stewart participated in several of these upgrading studies. He was elected as a Fellow of the Institute of Electrical and Electronics Engineers in 1987 for advances in transmission line theory and its reduction to practice through prototype demonstration. He is presently Chairman of the Transmission and Distribution Committee of the IEEE Power Engineering Society. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE.

3-1

Overview of Power Systems

Introduction
For railroad personnel to fully comprehend the issues involved in mitigating ac interference in their systems, they must have a basic understanding of power system design and operation. This section is provided to provide a minimal background in power system design and operation with emphasis on the aspects of the power system that most contribute to the railroads ac interference environment. A general overview of the power system is provided covering electric company generation, transmission, and distribution. Customer generation and operational issues will be included. Each aspect will then be reviewed in some detail, including which pieces of equipment and operational conditions affect the railroad environment. This information is summarized in tabular form in the quick-start version. Photographs will be used to assist in recognition of equipment and configurations. The reader should remember that electrical equipment for a given function could vary in appearances. Likewise, equipment with different functions can look very similar. The samples depicted in the photographs should help the reader identify potential issues, but the local power company personnel should always be consulted for confirmation of the types of equipment and issues present.

Quick-Start Version
Introduction to the Power System The electric power system consists of three main divisions: generation, transmission, and distribution. Some power companies also distinguish subtransmission as lying somewhere between transmission and distribution. Prominent characteristics of these areas of the power system are identified in the table below. The power system exists to supply uninterrupted electricity to its customers. As such, reliability considerations are paramount and all components (lines, transformers, substations) are continuously kept in service with a minimum of down time. Reliability criteria dictate how the system must perform in the event of various levels of contingency, such as the unanticipated loss of a transmission line segment. Lightning protection is an important part of system design for reliability and dictates grounding practices at both the transmission and distribution level. Other considerations such as personnel safety and protection of equipment also enter into the specification of grounding practices. Safety and fire protection considerations have led to secondary system grounding requirements to prevent damage or injury in the event of a cross between primaries and secondaries. For example, grounds are applied at each service entrance to limit voltages to ground at the customers locations. From the standpoint of electromagnetic compatibility between electric power and railroad facilities, conducted current into the earth from power system ground connections is the source of conducted interference. Sources of earth current and their characteristics are given in the second table below. Other EMC considerations include electric field coupling from the voltage 3-2

Overview of Power Systems

on overhead lines and magnetic field coupling from line current in both overhead and underground lines.
Table 3-1 Component Parts of the Power System Generation Transmission Converts sources of energy such as fuel, falling water, or nuclear material to electricity Bulk power transport connecting generating stations to substations serving load areas 69 765 kV Long lines with few, if any, taps or customer connections May be overhead or underground; underground circuits are short because of charging current limitations Loading limits for lines and loading areas including normal, long term emergency, and short term emergency limits Loading may be limited by system stability, voltage control, or conductor temperature Loading follows a cycle with time of day, season, and weather Not distinguished in many systems; lower level bulk power connecting transmission and distribution substations 23 115 kV May serve large loads directly Local connections to supply customers or groups of customers 4 35 kV Short lines with many taps, laterals, and branches; direct customer connections; may be reconfigured from time to time Loading is especially variable as customers turn devices on and off May be overhead or underground, with underground especially in urban or new residential areas

Subtransmission

Distribution

3-3

Overview of Power Systems

Table 3-2 Sources of Current Entering Earth from the Power System Transmission lines Structure grounds; structures connected together through shield wires Fault current Distribution lines Neutral grounds on multi-grounded primaries Secondary grounds at service entrances and distribution transformers Normal load current and fault current Current may be rich in harmonics Capacitor banks with blown fuses are a common source of harmonic interference Especially significant on single-phase portions of feeders Current variable as load changes Substations Substation ground mat is generally the lowest resistance path to earth With multiple distribution lines leaving a substation, the earth around the substation can be rich with neutral current

Detailed Version
Introduction to the Power System The interconnected electric power system is an integrated entity comprising generation, transmission, and distribution functions generally extending over a large land area. This system has evolved over more than 100 years. For many purposes it is reasonable to consider the generation, transmission, and distribution functions separately because they have different design requirements and constraints. For example, the transmission system is designed to efficiently transport bulk electric power long distances from generating stations to substations serving a given load area. Transmission lines are point-to-point affairs. The distribution system is designed to locally supply power to the end user. Distribution lines are complex, reaching down to each customer. As with all generalities, there are exceptions. Some large industrial customers receive power directly from the transmission system. Some small generators feed directly into the distribution

3-4

Overview of Power Systems

system. In addition to transmission and distribution, some utility companies also distinguish a subtransmission system intermediate between transmission and distribution. These transmission and distribution facilities represent a very large investment in fixed assets.

Figure 3-1 Overview of the Electric Power System

In North America the electric power system is predominantly 60 Hz ac with three-phase transmission and primary distribution. Europe uses predominantly 50 Hz ac systems. Some DC transmission lines exist, especially for long lines carrying heavy loads. The control possibilities that exist with DC, together with relative freedom from inductive and capacitive reactance, make it appealing for long-distance heavily-loaded applications where the cost of converter stations at

3-5

Overview of Power Systems

the ends of the line can be justified. In addition, some 25 Hz power remains, and in some places a few 2-phase customers remain. This discussion will focus on 60 Hz ac power systems. Voltage ranges for transmission and distribution vary by geographic location and company. There can be some overlap between what is considered transmission and what is considered distribution. Almost all transmission is three-phase, although some single-phase transmission exists in the Northeastern United States to supply the northeast corridor passenger railroad. Primary distribution lines are three-phase with single-phase laterals, especially in residential areas. In rural areas primary distribution lines can be three-phase, two-phase, or single-phase with increased distance from the distribution substation. Voltages are generally specified on a phase-to-phase basis. On a three-phase system the phase-to-phase voltage is 3 (1.73) times the phase-to-ground voltage. Transmission voltages in the United States generally range from 69 to 765 kV; subtransmission voltages generally range from 23 to 115 kV; primary distribution voltages generally range from 4.8 to 34.5 kV. The power system must reliably supply power to the customers with a minimum of interruption. As a result elaborate precautions are taken for reliability. The N-1 criterion demands that the overall power system continue to supply the entire load with any single element out of service. In order to attain this, power systems and their components typically have ratings for normal, long-term emergency, and short-term emergency conditions. Loss of a single transmission line cannot be allowed to push another line beyond its short-term emergency rating. Generation must be rescheduled or other changes made within the time defined by the short term emergency rating in order to keep the electrical system components operating within long term emergency ratings. The long-term emergency ratings may not be approached indefinitely. Various reliability councils, power pools, states, and individual companies have developed sets of reliability criteria necessary for proper operation of the system. A significant part of reliability centers on relaying and system protection. A faulted line section or other system component must be electrically isolated from the rest of the system in a 1 very short time . The required isolation time depends on whether the fault involves transmission or distribution. Relay protection involves sensitivity to detect faults, selectivity to identify the faulted object and not to falsely trip something that is not faulted, and sufficient speed to maintain stability and limit damage. Relay protection is part art and part science, and involves coordination of a number of different protective devices. On a distribution feeder the main substation circuit breaker must be coordinated with downstream fuses and reclosers (a.k.a. automatic reclose breakers) along the feeder to allow isolation of a fault with minimum disruption of service to customers. Transmission protection involves various levels of backup protection, so in the event of the failure of a relay to detect a fault or a circuit breaker to interrupt the fault, backup protection will operate to remove the fault, although at the expense of additional system components being tripped. Limitations on loading of transmission and distribution facilities are more complex than may be generally thought. One limit on power lines and equipment is their thermal capacity. Excess
A fault is a short circuit failure from one phase wire to another, or to ground. A relay is a device that senses a fault. A faulted line that has been electrically isolated is said to have tripped.
1

3-6

Overview of Power Systems

heat over time weakens metal and causes a variety of problems. The result can be structural failure (e.g. a broken wire) or reduced life (e.g. expected transformer life reduced from 30 years to 20 years). Normal thermal ratings are established such that no damage or reduced equipment life will result from continuous operation at these levels. Because thermal damage can be cumulative and is a function of temperature and time, long-term emergency ratings and shortterm emergency ratings are established. Thermal ratings are a function of season of the year and weather conditions. Thermal characteristics are different for different things; for example the thermal time constant of an underground cable is significantly longer than the thermal time constant of an overhead line. An overhead line achieves its new temperature in a few minutes, while an underground line takes many hours to reach a new equilibrium temperature. This makes a difference in how normal and emergency ratings are determined. Load cycles are factored into thermal ratings of underground cables, but not overhead line conductors. Another limit on power lines is their clearance from the ground or objects. These limits are a function of voltage and are established by various Codes, Standards, regulatory bodies, as well as good engineering practice. Reduced clearance generally results from increased line sag. Although, reduced clearance can also result from growth of trees, changes in the terrain, and construction. Worst-case clearance usually results from either high loading conditions (conductor heating) or heavy ice accumulation on the lines. Sometimes clearance limits are reached before thermal limits. In addition to load capacity considerations, there are voltage drop criteria to be met to ensure voltages remain within acceptable limits. Shunt capacitors and reactors are used for voltage control, as are transformer taps and adjustment of excitation of generating units. Active devices such as static var compensators are increasingly applied for voltage control. A power system may face voltage collapse under heavy loading conditions with insufficient sources of reactive power. Such voltage collapses have been the cause of blackouts. The transmission system also is faced with stability concerns, both steady state and transient. An analogy to power system transient stability can be made by considering a collection of weights suspended from a number of springs. The springs can hold a certain amount of weight if the load is applied gradually. However, if the load is applied suddenly, the springs may break. Likewise, because of the series inductive reactance of power transmission lines, the power system can handle a certain amount of load if the changes are gradual, but may become unstable for the same load if the changes are sudden. Lines have an electrical phase angle difference between ends, which is related to power flow on the line. If this angle exceeds 90 electrical degrees, attempts to increase the load increase the angle, but reduce the power flow on the line, an unstable condition. This is the technical background for many blackouts. A line trip resulting from a fault on a heavily loaded system can cause a redistribution of load among parallel paths. As the system swings to attempt to attain a new equilibrium operating condition, if the angles open too far, the system can become unstable. If the angle opens to 180 degrees during swinging, the condition is mathematically the same as a fault, the line protection operates, and an additional line trips. This further loads the rest of the system. A cascading failure may result. The overall power transmission system is very complex, with numerous lines, buses, and generators. Analysis of this level of complexity requires digital computer load flow, fault, and stability programs to study conditions over a widely ranging area. Different contingencies must 3-7

Overview of Power Systems

be evaluated to ensure that they do not result in instability, voltage collapse, or pushing other lines or equipment beyond their short-term emergency ratings. Historically, the generating and transmission systems were designed together to make an integrated whole. Operating procedures were developed to handle different types of contingencies. For example, sometimes relaying would be designed so that under certain conditions a line trip would also transfer trip2 a generating unit to preserve stability. This whole area of system engineering has been complicated by deregulation, where the generation siting and transmission planning functions have become decoupled. Reliable operation of the transmission system is an increasingly difficult challenge. As a result of the constraints of thermal limitations, voltage control requirements, and stability, the power system is required to operate with all facilities intact as much as possible. Maintenance outages are few and far between, especially on heavily loaded portions of the transmission system. Outages are also few on the distribution system, especially where customer outages may also be involved. Other facilities near power transmission and distribution lines must therefore adapt to their continuous presence. Transmission and distribution facilities may be overhead or underground. Overhead construction is less expensive, and more subject to disruption by weather, but more quickly repaired than underground construction. Outages on the underground system tend to be less frequent, but of longer duration, than on the overhead system. Underground systems have different electrical characteristics from overhead systems due to the greater line capacitance in underground systems. Combined overhead and underground lines are becoming more common. For example, it may be impossible to secure a right-of-way for a portion of an overhead transmission line route, making it necessary to construct a portion of the line underground. Overhead Transmission Lines Overhead transmission lines consist of a number of components: foundations, structures, conductors, insulators, various types of hardware (e.g. to connect the insulators to the structures and the insulators to the conductors), and overhead shield wires. Electrical, mechanical, and civil engineering all are involved in design of overhead transmission lines. Electrical concerns include insulation for power frequency voltage, switching surges, and lightning, as well as electrical environmental effects such as radio and television interference, audible noise, and electric and magnetic field coupling to nearby objects. The design of structures and foundations is a civil engineering function. Mechanical engineering design variables include such things as dampers to eliminate conductor damage due to aeolian vibration in wind, ice galloping, and thermal ratings of conductors. Thermal rating is more of a mechanical than an electrical engineering issue because it is primarily a heat transfer phenomenon. Transmission line structures come in a number of different types: wood pole, steel pole, concrete pole, and steel lattice. They can be self-supporting or guyed. A single line can have many different structure configurations. Mechanical loading requirements differ widely depending on several factors. Among those factors is whether the structure is a tangent (straight line) structure, light angle structure, or
A transfer trip is when a relay senses a fault at one location and, through high speed communications, commands fault clearing switches at another location to trip.
2

3-8

Overview of Power Systems

heavy angle/dead end structure. Individual transmission line structures may support one, two, or more three-phase circuits in addition to the shield wires. Lower voltage transmission or distribution circuits may be constructed underneath the main circuit (underbuild).

Figure 3-2 Overhead Transmission Structure Types

The overhead shield wires are installed above the phase conductors and provide the primary lightning protection for transmission lines. Lightning failures of transmission lines result from two mechanisms, called shielding failures and backflashovers. A shielding failure occurs when lighting is not intercepted by the shield wires and directly strikes a phase conductor. This tends to happen for lightning flashes having relatively low stroke current. The number of shield wires and their placement relative to the phase conductors are crucial to achieving acceptable shielding failure performance. An additional function of the shield wire is to provide a return path for power frequency fault currents. The shield wire must be sized to handle the maximum fault 3-9

Overview of Power Systems

current it might be called upon to carry. The shield wire reduces the zero sequence reactance of the circuit, increasing the fault current. The effect of the shield wires on fault current flow is desired by system protection engineers for enhanced relay protection. Lightning is a phenomenon best considered as a current injection of thousands of amperes in a microsecond time frame. Because electromagnetic waves travel approximately 300 meters per microsecond, for a lightning impulse cresting in 1.5 microseconds, the leading edge of the current wave on the shield wire has only traveled about 450 meters. Thus, lightning performance is a very localized phenomenon as opposed to power frequency insulation, which covers the entire line. A backflashover occurs when the lightning-caused voltage rise on a structure is sufficiently above the impulse strength of the insulation so that the voltage flashes over from the structure to the conductor. This is backwards from a shielding failure where the conductor rises to a higher voltage than the structure. Perhaps the most important single parameter for preventing lightning backflashover is reducing the footing resistance to earth of the structures. For acceptable lightning performance, especially in high lightning activity areas, it is essential that the shield wires be grounded to earth through low resistance grounds at each structure. There is insufficient time for the traveling wave to travel down the line to find a good ground to prevent flashovers, making it necessary for each structure to be well grounded. As part of the lightning protection, wood pole structures have down leads connecting the shield wires to grounded electrodes (ground rods or butt wraps) at the base of the poles. There are two approaches to grounding the point where the insulators connect to wooden structures. Sometimes the connection points are connected to the down leads to firmly ground the structure end of the insulator. Other designs use the wood as part of the impulse insulation strength with the down lead spaced away from the poles for a distance near the insulator attachment points. However, leakage current in the wood in this arrangement is a potential cause of pole fires. In addition to lightning, transmission lines must be designed for adequate power frequency voltage and switching surge overvoltage conditions. Power frequency voltage insulation design is especially related to the contamination performance of the insulators, which is in turn related to the type of insulator and its leakage distance. Leakage distance is one of the insulator characteristics presented in insulator catalog data. Contamination performance and leakage distance considerations are the reason insulator strings are longer for the same voltage transmission line in areas such as seacoasts with extensive salt contamination. From the standpoint of electromagnetic compatibility, the transmission line can be considered as a set of conductors equal to the total number of phase and shield wires. For example, a double circuit three-phase line with 2 shield wires would have 8 total conductors, Figure 3-9. Transmission lines are frequently considered to be well balanced, that is, three voltages of equal magnitude spaced 120 electrical degrees and three currents of equal magnitude spaced 120 electrical degrees. For balanced conditions the three phase voltages sum to zero and the three phase currents sum to zero. Thus, for balanced conditions there is no current returning anywhere other than in the phase conductors. But in reality, there is always a small amount of unbalance in the phase currents, and occasionally the unbalance can become a significant percentage of the average phase current. When the three phase currents do not sum to zero3,
This is typical in the transmission system because neutrals are grounded at Y-connected transformers in substations.
3

3-10

Overview of Power Systems

some of the unbalance current returns in the shield wires and some returns in the earth. Some of this current flows between earth and the shield wires at the structure grounds because of the relative impedance differences of the different paths.

Figure 3-3 Double Circuit Line with 2 Shield Wires 8 Total Conductors

Current also flows in the shield wires as a result of magnetic field coupling from the phase conductors. Current may also flow in the shield wires for other reasons. A transmission line connects substations that may be separated by a number of miles. As the potential of the ground mats of the substations may be different, a current may flow in the shield wires as a result of the connection they create between the substation ground mats. Sometimes attempts are made to segment shield wires to reduce losses or for other reasons, but the drawbacks may preclude using anything other than a continuous shield wire. In addition to normal load current, currents on the order of tens of thousands of amperes flow for short periods during faults. It is necessary to isolate and eliminate faults within a few cycles4, both for stability reasons and to limit arc damage. Thus, transmission line fault currents are short-lived. However, for the fault duration, currents significantly greater than load current

One cycle lasts 1/60 of a second at 60 Hz. Fault clearing times on transmission typically last 2 to 6 cycles.

3-11

Overview of Power Systems

occur. When faults from phase to ground occur, the fault current returns both in the earth and the shield wires. The division of current between earth and shield wires depends on the relative impedance of these two parallel paths. Both near the fault and near the substation, substantial currents flow in the structure grounds as the fault current seeks the proper division between earth and shield wires. Step and touch potentials are concerns at structures and substation fences during faults. Underground Transmission Lines Underground transmission lines fall into three main categories: High Pressure Fluid Filled (HPFF) Cable: These are paper-insulated cables installed in steel pipes filled with insulating fluid maintained under high pressure. Three phases are installed in a single pipe. This is the predominant type of transmission cable in old installations in North America, although it is rarely installed today. Of the three cable types, construction of three phases in a relatively small single steel pipe results in low external magnetic fields and the least induction to parallel objects. Low Pressure Fluid Filled Cable (a.k.a. Self-Contained Fluid Filled Cable): These are cables where the insulating fluid is at relatively low pressure internal to the phase cables. They may be installed in steel or plastic pipes. This type is typical of 345 kV cables being installed today. Cross-linked polyethylene (XLPE) (a.k.a. Solid Dielectric Cable or Extruded Dielectric Cable): These are cables with solid insulation, and are typical of lower transmission voltages. Occasionally direct buried, these cables may be installed in steel or plastic pipes.

Each conductor of an underground cable has a construction consisting of the central core conductor, a semiconducting covering, the main cable insulation, a second semiconducting covering, a conducting sheath, and the protective outer covering. The outer protective covering may be lead, a plastic material, or in the case of submarine cables a wrap of armor wires. The grounded conducting sheath is necessary to insure a uniform electric field gradient within the cable insulation and prevent locally high fields that would degrade the insulation and cause electrical failure. Grounding of the cable sheaths has an effect on circulating current in the sheaths and on the magnetic field outside the cables. Multi-grounded sheaths have the sheaths grounded at least at the two ends of the cable. A circulating current arises in the sheath because of magnetic induction between the central conductor and the sheath. The magnetic field from the circulating current partially cancels the magnetic field from the phase current, lowering the external magnetic field. The circulating current and resulting magnetic field cancellation are functions of the electrical resistance of the sheath. The circulating current in the sheath results in a power loss in the sheath, generating heat and reducing the cable ampacity. In order to reduce the power loss and increase cable ampacity, the sheaths may be single point grounded, that is, grounded at only one end. This preserves the function of the sheath in grading the electrical stress in the cable insulation, but eliminates the circulating current. Power loss is reduced, but external magnetic field is increased. A third option is cross-bonding, where the ground path may go from one end on the sheath of phase A to the sheath of phase B at approximately the 1/3 point along the cable, 3-12

Overview of Power Systems

to the sheath of phase C at approximately the 2/3 point along the cable. The induced voltages along the three segments approximately sum to zero, making the sheath current small. This preserves the redundancy in grounding of multi-grounded sheaths while virtually eliminating sheath currents. As with overhead lines, it is necessary to consider fault currents and the paths they follow. If the cables are installed in a steel pipe, the steel pipe itself is the primary return path for fault current to ground. Overhead transmission lines have widely spaced conductors with air acting as the insulating dielectric. Underground transmission lines have closely spaced conductors with solid dielectric material with a higher dielectric constant. Thus, underground lines have greater capacitance per mile than overhead lines. Current flowing into the line capacitance is called charging current. Because of this charging current, underground transmission systems have some fundamental differences from overhead systems. At higher voltages, the charging current limits how long a cable circuit may be before charging current uses up the entire cable ampacity. Charging current also is a concern for voltage control, not only during normal operation, but especially during abnormal conditions such a restoration from a power failure. Careful coordination of shunt reactors may be required to limit overvoltages. This is especially important because if cable insulation is damaged, a long and costly repair is required. In contrast, an overhead line can withstand flashovers without damage. Overhead Distribution Lines Whereas transmission lines are generally point-to-point (although there are occasional threeterminal lines), distribution lines have rather complex topology. A distribution line starts at the distribution substation, may proceed for several miles along what is termed an express feeder, and then serve loads through many branches, taps, and laterals. Distribution engineers refer to the high voltage side of the distribution line as the primary and the low voltage connection to the customer as the secondary. For example, a three-phase 12,470 volt feeder may start at a substation; a 7200 volt single-phase tap can branch off down a side street; and a single-phase 7200 to 120/240 volt transformer would provide 120/240 volt secondary service to a group of houses.

3-13

Overview of Power Systems

Figure 3-4 Distribution Transformer Feeding Single-Phase 120/240V Load from a Three-Phase wye-Connected 12,470/7200V Primary with a Multi-Grounded Neutral

Distribution voltages and grounding techniques are dependent on local history and practices. There may also be a mixture of voltages and grounding practices on a single feeder. For example in some locations it was common years ago to install 4800-volt delta-connected feeders with transformers connected line-to-line. Over a period of time these feeders may have been upgraded, perhaps to 12,470-volt wye-connected multi-grounded lines. For cost or other reasons, it may not have been practical or economic to convert the entire feeder. Isolated sections may remain 4800-volt delta fed from the 12,470-volt feeder through transformers called ratio banks by the line crews. Construction practices differ with locality and operating company. Poles may be wood, steel, or reinforced concrete. Wires may be supported on pin-type insulators mounted on crossarms, they may be on pin-type insulators on candlestick brackets, or on post insulators mounted directly on the poles, among other possible configurations. The same poles used for power distribution lines are also frequently used for telephone and cable television wires, necessitating joint use agreements. Messenger wires for telephone and cable television cables are metallic conducting paths parallel to the distribution circuit neutral and carry some of the neutral current.

3-14

Overview of Power Systems

Figure 3-5 Distribution Pole with Telephone and Cable Television Wires

Primary Feeder Grounding Distribution primary feeder grounding practices are different in North America from what they are in Europe. In the United States few customers are fed from a single transformer, except for urban secondary network systems. So, the length of secondary wires is limited and the primaries are relatively long with branches and taps giving significant complexity. The general practice in Europe is to feed many customers from a single secondary of considerable physical extent. Different secondary voltages (120 volts in North America, 240 volts in Europe) contribute to these differences in design philosophy, which then lead to differences in feeder grounding practices. Different practices prevail in different parts of North America as well, with different philosophies often present in the same geographic area, depending on the age of the feeder. As far as grounding is concerned, primary feeders can be described under several broad categories: 4-wire multi-grounded. 4-wire single point grounded 3-wire ungrounded 3-wire grounded (unigrounded) 5-wire 3-15

Overview of Power Systems

4-Wire Multi-Grounded Feeders

In North America, 4-wire multi-grounded feeders are the most common choice today for cost and safety reasons. A three-phase multi-grounded feeder consists of 4 wires: 3 phase conductors and a neutral wire often mounted on the pole below the phase wires. One phase and the neutral supply single-phase loads. The ease and low cost of supplying single-phase loads is a major reason for use of the 4 wire multi-grounded system in the United States. Most customers in the United States are single phase, and more than half of a typical distribution feeder is single phase. Transformers are connected phase-to-neutral in this arrangement, allowing the use of single bushing transformers with a single surge arrester and fuse per transformer. The neutral wire in parallel with the earth forms the path for unbalance currents to return to the source at the substation, both during normal operation and during faults. The neutral wire in parallel with the earth forms a low impedance path for fault current back to the source. Because the neutral wire is in parallel with the earth, the earth serves as a return conductor for part of the unbalance current. Typically the division is assumed to be approximately 50% in the neutral wire and 50% in the earth.

Figure 3-6 4-Wire Multi-Grounded Feeders

The neutral wire in this arrangement is grounded at the substation, at transformers, and possibly elsewhere. Ground connections are frequently made to the neutral wire at least every 500 to 1000 5 feet along a feeder, with code requirements of a minimum of four grounds per mile . Lightning arresters are also grounded to the neutral wire. A low resistance path from neutral to earth still exists in the event of a failure of a single ground connection. Even if the neutral wire is severed, a neutral path still exists to the substation through the various connections to earth. A very important advantage of the multi-grounded neutral design is the significant safety benefits it provides both for customers and power company crews. System fault protection can more easily clear faults with a multi-grounded neutral. The neutral acts as a physical barrier, as
In the United States, very old feeders built under earlier versions of the NESC (Code) might not have four grounds per mile.
5

3-16

Overview of Power Systems

well as providing a solid fault current path when a phase conductor contacts the neutral wire. Grounding the neutral helps prevent dangerous step and touch voltages during such faults. Another significant advantage to the multi-grounded feeder is limitation of neutral shift and local voltage rise between phase conductors and ground, especially under fault conditions. On a threephase feeder, during a single-phase fault to ground, the local voltage to ground rises on the other two phases6. The insulators and lightning arresters selected are a function of the amount of voltage rise possible on the un-faulted phases. Ungrounded systems require insulators and arresters that are rated for at least the full line-to-line voltage. Multi-grounded systems allow the use of insulators and arresters with lower voltage ratings (but still at least line-to-ground voltage). One consequence of the use of the multi-grounded neutral is that current flows into the earth at the neutral wire grounding locations, both during normal operation and during faults. While the current flow into the earth is much greater during faults, it still may be significant during normal operation.
4-Wire Single-Point Grounded Feeders

The 4-wire single-point-grounded system also consists of three phase wires and a neutral wire. The difference is that the neutral wire is grounded only at the substation. This arrangement has been used in special situations. Single-point neutral grounding presents a problem for grounding of lightning arresters, because they require a good low resistance local ground connection for proper operation. A concern with the single-point grounded system is the safety issue if the neutral wire were to break or become open7. In a single-point grounded system, if the neutral wire is open, the path to the substation is severed. Since distribution transformers are connected line-to-neutral, there is the possibility of significant overvoltages with an open neutral wire if there is only one ground connection located back at the source. On the other hand, a benefit of the 4-wire single-point grounded system is that there is no return current injected into the earth during normal operation.
3-Wire Ungrounded Feeders

The original premise of the ungrounded feeder fed from a delta-connected transformer was the ability for the feeder to continue to operate in the presence of a single fault between a phase and ground. Many feeders of this type were installed in the North America, especially from the 1920s through the 1940s. The presence of a fault would be identified by a change in the three phases voltages to ground, and time would be allowed for location and repair of the fault. The prevailing theory was that during a single phase fault to ground the other phases would rise to line-to-line

The amount of voltage rise on un-faulted phases is related to the ratio between zero sequence equivalent reactance and positive sequence equivalent reactance of the system at the fault location. The multi-grounded configuration has lower equivalent zero sequence reactance and lower voltage rise on the un-faulted phases. In contrast - in a multi-grounded system, the neutral path to the substation is preserved through the various ground connections and the earth even if the neutral wire is open.
7

3-17

Overview of Power Systems

voltage above local ground. This requires insulators and arresters rated for line-to-line voltage rather than line-to-ground voltage. Lightning arresters still require local low resistance grounds.

Figure 3-7 3-Wire Ungrounded Feeders

The problem with this approach is that in reality, all systems are grounded through parasitic capacitance in the absence of a resistive connection to earth. Thus under certain circumstances, it is possible to have the line-to-ground capacitance resonate with the system inductance. Especially during a fault condition a series resonance may occur between capacitance and the feeder and/or transformer reactance. This condition results in a voltage on the unfaulted phases possibly considerably in excess of phase-to-ground voltage with consequent possibility of insulation damage, fire, or personnel injury. Arcing ground faults could also excite resonances and result in very high overvoltages and consequent damage. The same reasoning was applied in industrial installations that had ungrounded wiring. As ungrounded delta feeders have been upgraded by increasing voltage, they generally have been converted from ungrounded feeders to multi-grounded feeders. Many times the cost of converting the entire feeder to higher voltage was not justified because of the amount of load, length of circuit, or other reasons. For example, an isolated housing development may not have been converted because of the need to change all the pole transformers. Because of this, many multi-grounded feeders continue to have ungrounded portions fed from transformers

3-18

Overview of Power Systems

(called ratio banks by linemen) on pole pedestals. It is not uncommon to have multi-grounded and ungrounded feeder segments in the same geographic area. 3-wire ungrounded systems are less expensive than 4-wire feeders for supply of 3-phase loads. The need to string 3 rather than 4 wires is a cost saving that led to the use of 3-wire systems in the early years of the power industry. Ungrounded systems have no return path through the earth, so no current flows in the earth during normal operation. This reduces magnetic field and conductive and magnetic field induced interference. However, some current can flow in the earth during faults to ground. The use of ungrounded systems has been declining for many decades in the United States both in primary distribution feeders and industrial installations. Many industrial systems have been converted from ungrounded to grounded systems by installation of grounding transformers. A zig-zag connection was developed specifically for that purpose, and has been widely used in industrial applications. Grounding transformers have also been applied to distribution feeders. When a grounding transformer is used, care must be taken to ensure that the grounding transformer does not become disconnected and allow the neutral to float (the zero sequence path must be closed).
3-Wire Grounded (Unigrounded) Feeders

A 3-wire grounded feeder (also called unigrounded system) is like the 4-wire single point grounded feeder, except that the neutral conductor is not taken out of the substation. It is similar to the 3-wire ungrounded feeder in that there are only 3 phase wires to the feeder. This has the advantage of holding feeder voltages closer to ground, but loses the advantage of being able to operate in the presence of a single fault to ground. This connection has the disadvantage that single-phase laterals require blowing of two fuses to clear faults. Blowing of only one fuse can subject customers to extended periods of abnormally low voltage. The 3-wire unigrounded system is more common than the multi-grounded system in California and much of the rest of the world outside the United States. Its use in California results from Californias rules for overhead line construction that require overhead neutral conductors be mounted on insulators and that the grounding conductors for the primary neutral be separate from those of the secondary neutral. Most of the unigrounded systems in California have solidly grounded neutrals, although some have low impedance reactors to limit fault current. In contrast to North American practice, European distribution systems are normally 3-wire with all loads being three-phase or connected line-to-line. European practice is to serve a larger number of customers per transformer than in North America. Most European transformers are three-phase and much larger MVA than the mostly single-phase American transformers. As a result, European practice has no need for a primary circuit neutral. Some European countries have restrictions on the use of the earth as a parallel return path based on protection of communication circuits from interference. This discourages use of multi-grounded neutrals. European residential customers may be supplied with 2 phases of a 3-phase secondary, as opposed to the United States practice of serving residential customers with two 120-volt connections 180 degrees out of phase. The European standard secondary voltage ranges from 220 to 240 volts, twice the 120 volts used in North America. With the higher secondary voltage, 3-19

Overview of Power Systems

European designs frequently have a multi-grounded secondary neutral of considerable extent with many houses on each transformer. The secondary grounding provides the safety grounding in this arrangement. With longer and more elaborate secondaries, European primaries generally do not have as many taps as in North America. Both 3 and 4-wire systems are used outside of Europe and North America, with European practices more widely followed. In some parts of the world a combination of the two is used. Perhaps the worst is where 120-volt secondaries are combined with European style primaries. Such a system combines the limited distance possible at 120 volts with acceptable voltage drop with the more expensive primary configuration.
5-Wire Feeders

The 5-wire design is a new approach developed as an attempt to reduce stray voltage and magnetic fields and to provide a detection mechanism for high impedance faults. The 5-wire design has 3 phase conductors, an isolated neutral grounded at only the substation, and a multigrounded ground wire. This is similar to a 3-phase building branch circuit, which has 3 hot wires, a neutral, and a safety ground. Transformers are connected between phase wires and the neutral. Lightning arresters are grounded to the multi-grounded ground wire. The isolated neutral carries the unbalance return current back to the substation isolated from earth. The multigrounded ground wire performs the safety functions of the neutral in the multi-grounded feeder. This approach has been demonstrated in a prototype project, but has yet to be used for an actual application.
Other Systems

A single wire earth return primary is used in some remote and rural parts of the world. In an earth return system the earth carries all the return current, and the resistance of ground connections is critical. This is an inexpensive approach for use in sparsely settled areas with long feeder distances and little load. However, voltage control and safety are issues. This was a common practice in some rural locations in Canada, where grounded neutrals were reestablished at each step-down transformer. The American National Electrical Safety Code (IEEE C2-1997) prohibits the use of the earth as the sole conductor for any part of the circuit, so this arrangement is not used in the United States.

3-20

Overview of Power Systems

Figure 3-8 Single Wire Earth Return Primary

High-resistance or high-reactance grounding is rarely used on distribution circuits in North America. European systems sometimes use a 3-wire resonant grounded system where the substation transformer is wye connected, with the neutral grounded through a reactor called a Peterson coil. The reactor is tuned to cancel the line-to-ground capacitance, forming essentially an ungrounded system. As an ungrounded system, it can continue to operate with a single phase faulted to ground. By tuning out the capacitance, fault current is reduced to the point where transient faults clear themselves. In practice it has often been difficult to keep the Peterson coil properly tuned to the system capacitance. Secondary Feeder Grounding The typical residential service in North America is 120/240-volt single phase with the neutral point grounded. Other services are available to commercial and industrial customers, a very common one being 120/208-volt 3-phase grounded wye. Delta service is occasionally found with a corner of the delta grounded, and sometimes a delta service with the midpoint of one side of the delta grounded. This latter is called a red leg delta system. The National Electrical Safety Code prescribes primary feeder grounding for the United States. Secondary feeders and building service entrances are covered by the National Electrical Code in the United States, and in Canada by the Canadian Electric Code Part 1. The National Electrical Code requires that service entrances be grounded to a metallic water pipe plus an additional 3-21

Overview of Power Systems

ground such as a driven rod. Transformer secondary neutrals are grounded. If there is a primary neutral, primary and secondary neutrals are connected together. The impact of this is to make a number of additional locations where the power system is connected to earth and consequently where current can enter the earth. Other Considerations Related to Ground Currents An additional consideration regarding neutral and ground currents in three-phase systems results from current harmonics. The positive sequence currents sum to zero in a balanced 3-phase system, leaving no current to flow in the neutral path, Figure 3-9. Likewise, any negative sequence currents in the phase wires resulting from unbalance sum to zero, leaving no neutral path current. In contrast, the zero sequence current returns entirely in the neutral path. 3N harmonics (in other words, 3 times any integer times power frequency, 180 Hz for example) are zero sequence harmonics whose current returns in the neutral path. With the proliferation of electronic devices such as adjustable speed motor drives, switching power supplies in computers and other electronic equipment, electronic fluorescent lamp ballasts, and similar items, harmonic currents have been increasing. With the increase in harmonic currents has come an increase in neutral wire currents and currents in ground leads and consequently in the earth. As a result, the waveform and harmonic current of neutral and ground currents can be quite different from those of phase currents because of the zero sequence effect, Figure 3-10.

Phase and Neutral Currents - 0% Harmonics


1.5 1 0.5 0 -0.5 -1 -1.5 0 100 200 300 400 A Phase B Phase C Phase Neutral

Figure 3-9 Positive Sequence Currents Sum to Zero in a Balanced 3-Phase System

3-22

Overview of Power Systems

Phase and Neutral Currents - 3% Harmonics


1.5 1 0.5 0 -0.5 -1 -1.5 0 100 200 300 400 A Phase B Phase C Phase Neutral

Figure 3-10 Zero Sequence Harmonic Current Returns in the Neutral Path

Another source of occasional ground currents results from high impedance faults on overhead distribution feeders. If a phase wire contacts a tree branch, or falls on an asphalt road surface, among other conditions, the fault current may be less than normal load current, and thus below the level needed to operate the feeder protection. Such a condition can persist for some time before detection and repair, and provides another possible point of ingress of current into the earth. Impact on EMC Current in the earth affects nearby railroad facilities in two ways. Current in the earth conductively couples to the railroad facilities grounds, as described in Chapter 2. Also, when part of the current returns in the earth, the current loop is enlarged, as compared to when all of the current returns in wires on the poles, thus increasing magnetic field strengths and consequent magnetic field induction. Therefore, distribution feeder designs that minimize current in the earth reduce possible interference effects. However, multi-grounded primary feeders and grounded secondaries have important safety advantages that drive their use. The 5-wire design is an attempt to have the best of both worlds, but remains to be proved in practice. Substations Near Railroad Tracks Distribution neutral current returns to the transformer that is its source. For this reason the current density in the earth from neutral return current is usually at its highest near the distribution substation. Because some of this energy can couple into the track circuit conductively, ac interference in the track may peak near adjacent substations.

3-23

Overview of Power Systems

Underground Distribution Lines Underground distribution lines are constructed of solid dielectric cables, either direct buried or in pipes or duct banks. Each phase has a central core phase conductor and a concentric neutral conductor operated in a multi-grounded fashion. As with overhead construction, there are threephase primary feeders with many branches, including single-phase taps. Pad mounted switches and transformers may be utilized.

Figure 3-11 Pad Mounted Transformers

The most common underground feeder configuration is the radial feeder, the same general layout as with overhead feeders. Such a feeder has a single source point, and branches out to supply the loads. A fault that takes out a cable takes out the feeder, except for the area between an intermediate coordinated fuse and the substation. Where high reliability is required, as in city centers, the network system has often been used. A network system is fed from several sources through several transformers with an interconnected mesh of secondary connections such that the loss of any circuit element would allow all the customers to continue to be served. While complex and expensive, the network system has found wide use in urban areas. Spot networks are often used for industrial and commercial loads. It is increasingly common to see combination overhead and underground feeders where a portion of the feeder is overhead and a portion is underground, thus combining the characteristics of both in one feeder. Pole mounted potheads make the transition straightforward.

3-24

Overview of Power Systems

Figure 3-12 Pole Mounted Potheads for Overhead/Underground Transitions

Underground feeders are less subject to damage from weather than overhead feeders, but damage from lightning is still possible. Lightning arresters are still required. Dig-ins and similar occurrences damage cables. Locating the resulting fault can be a challenge. Secondary Distribution Secondary distribution concerns the low voltage connections provided to the great majority of customers. Residences in North America are usually supplied 120/240 volts single phase. Many commercial customers take service at 120/208 volts three-phase. Other voltages are sometimes employed, for example 480 volt three-phase commercial customers. Some large industrial and commercial customers are directly connected to the primary feeder and are said to take primary service. Because of the low voltage employed by residential and many other customers, secondary connections are generally short, except possibly in network systems. Secondary grounding practices in the United States are specified in the National Electrical Code (NEC) for reasons of fire protection and human safety that go back 100 years. The code requires that customers service entrances be grounded, usually to a water pipe plus an additional ground rod. In addition, at the transformer, the primary and secondary neutrals are to be bonded together in most cases. Bonding the primary and secondary neutrals has advantages and disadvantages. While this practice can allow primary neutral current to flow into the secondary customer neutral wire, it also limits the voltage rise that can occur if the primary and secondary phase wires accidentally become interconnected, for example because of a downed primary wire falling across a secondary wire. 3-25

4
OVERVIEW OF RAILROAD SIGNAL CIRCUITS
This chapter provides a comprehensive explanation of railroad signal systems, with concentration on the issues relevant for electromagnetic compatibility. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA. Bennett R. Feely has been engaged in the railway signaling profession throughout his career, beginning in Philadelphia, PA, as a signal maintainer on the subway-elevated lines of the Southeastern Pennsylvania Transportation Authority. He has since worked as a signal engineer for General Railway Signal and Union Switch & Signal throughout the past 25 years, specializing in wayside devices, circuits and systems. Areas of special interest include track circuits, switch machines, relays and light signals. Through his work and studies he has gained detailed insight and expertise in technical and commercial matters pertaining to the design, manufacture, installation and maintenance of railway signaling equipment. Mr. Feelys broad range of experience also includes participation in important projects such as NYCTs Canarsie CBTC project and Amtraks Northeast Corridor Improvement project. Mr. Feely attended the University of Pittsburgh and Penn Technical Institute. He holds several patents in the field of railway signaling.

4-1

Overview of Railroad Signal Circuits

Introduction
This chapter is designed to provide the reader with a little of the needed familiarity with railroad signaling equipment. Once the specific equipment suffering from an apparent case of ac interference is identified, the detailed version of this chapter may serve as reference material on that equipment. Otherwise, the chapter may be used as a compendium of railroad signaling history and background information from which the reader will be able to draw some insight into related or similar railroad signaling equipment.

Quick-Start Version
The current state of modern railroad signaling is the evolutionary result of decades of history. Yet the historic elements remain relevant, because there is a great deal of older signaling equipment still in use. This legacy equipment also serves as a springboard for the development of new ideas. The most common types of railroad signaling equipment usually fall into one (or more) of the following categories: 1. systems designed to detect the presence of a train within an area defined by a track circuit transmitter, and a track circuit receiver (e.g. dc track circuits, AFO track circuits, etc.) 2. systems designed to communicate information (such as a trains location) along a railroad line (e.g. coded track circuits, line-wire circuits, radio communications) 3. systems designed to measure a trains position or motion with respect to a fixed point (e.g. Motion Sensors, Crossing Predictors) 4. systems designed to detect specific hazards to railroad operations (e.g. slide fences, dragging equipment detectors, etc.) 5. systems designed to provide safety-critical information to trains or motorists (e.g. wayside signals, cab signals, crossing flashers, crossing gates, bells) 6. systems designed to physically reconfigure the railroad tracks to construct a particular route of travel for a train (e.g. switch machines, switch heaters, switch locks, etc.) Although most of the types of equipment listed above can suffer from ac interference due to nearby ac power transmission and distribution systems, and some can even help create it, the first indication of an ac interference problem will usually come from the Motion Sensors and Crossing Predictors used to control the warning devices at grade crossings. They are often the most sensitive detectors of unwanted ac electrical energy on railroad tracks. Therefore, particular attention should be paid to these systems.

4-2

Overview of Railroad Signal Circuits

Having an understanding of the specific use of a piece of railroad signal equipment, and how it performs this function, is key to understanding its susceptibility to ac interference. In general, unwanted ac energy can get onto railroad tracks by magnetic induction, electric induction, or ordinary conduction (via both intentional and unintended paths). This ac energy can prevent signaling equipment from working properly by either preventing the proper transmission or reception of the signaling currents, voltages, and frequencies used on the railroad (think signalto-noise ratio), or by permanently damaging the signaling equipment. When coming to grips with any unfamiliar piece of railroad signal equipment, the most important questions to keep in mind in this context are: 1. What specific function does the affected piece of railroad signal equipment perform? 2. How does it perform this function? 3. How can the presence of unwanted ac power-line-related voltages or currents affect or impair the normal operation of this equipment? 4. How susceptible to ac interference is the affected piece of railroad signal equipment? 5. How can the normal performance of this specific signaling function be restored in the presence of unwanted ac voltages or currents? 6. What alternative technologies or equipment types could be used to avoid this problem? Identifying the functions and operating characteristics of a piece of railroad signaling equipment, and finding the answers to the questions above, are the essential foundations upon which an investigation and resolution of ac interference problems must be based.

Detailed Version
Railroad Signaling Circuits Introduction Railroad signaling is the branch of Electrical Engineering concerned with the electric and electronic systems designed to promote the safe and efficient movement of trains. Although historically focused on equipment using Track Circuits, this field continues to expand, and now includes radios, high-speed digital communications, and modern computing technologies. Railroad signal systems are required in the United States by the rules of the Federal Railroad Administration (FRA) for some railroad operations, and are an economic necessity on many heavily traveled lines. While the primary purpose of early signaling systems was improved safety, it was quickly recognized that signaling technology could also substantially improve a railroads productivity, and hence its profitability. A real or expected increase in train traffic is often the motivation for railroads to install signal systems. But once installed, a signal system must be maintained in good working order, and remain essentially unchanged, unless the proper paperwork has been filed with, and approved by, the FRA. 4-3

Overview of Railroad Signal Circuits

The technology of railway signaling is well established, but has been continuously changing and improving since the earliest beginnings of railroad signaling in the 1870s. Although the authors have tried to restrict this chapter to only those signaling systems that are currently in use, this narrows the field by very little, as there is a great deal of older signal equipment still in use. Due to the rapid pace of advances made in railroad signaling over the past several decades, and the longevity of the electrical, electromechanical, and electronic devices used, current railroad signal systems frequently have brand-new microprocessor-based equipment working side-by-side with relays, wires, and other components that have been in continuous service for more than fifty years. In this section, we will begin by discussing the various mechanical parts of railroad tracks and railroad cars that are used by the signaling circuits to detect the location or presence of a train. This will be followed by a review of most of the various types of railroad signaling equipment currently in use in North America, as well as a few systems used elsewhere. Where appropriate, we will discuss the Electro-Magnetic Compatibility (EMC) issues associated with each type of equipment as it is introduced. There is also a brief summary discussion of signal equipment EMC issues at the end of the chapter. The hazards arising from EMC issues are especially worrisome to the signal maintainer, because the cause and effect relationship is often transient and thus difficult to observe and analyze. Fortunately, the fail-safe nature of the equipment assures that the vast majority of failures will be so-called right-side failures, in which the signal system takes appropriate action to protect the safety of trains and the public. Railroad Track Part of the Track Circuit Unlike most other surface vehicles, trains are not steered by their operator or pilot. The engineer has control of the trains throttle and brake, but cannot control its direction of travel beyond the selection of forward or reverse. As with monorails, roller coasters, and industrial conveyor systems, it is the track itself that steers the train. As such, railroad tracks are often viewed as purely mechanical components. But since the time of Robinsons track circuit patent in 1872, the rails of the track have also been the most essential electrical components of railroad signaling. The most novel part of Robinsons patent was that it used the electrical transmission characteristics of the track itself to detect the presence of a train. An electrical current was transmitted down the track using both rails. Thus, the rails of the track became part of an electrical circuit known as a track circuit. This basic principle has been a fundamental part of railroad signaling practice for the past 130 years in North America. Track circuits of one type or another can be found on almost any railroad in North America. However, many components of these track circuits are hidden away in the equipment enclosures located along the railroad right-of-way. At its most basic, a track circuit consists of: 1. a source of electrical energy, such as a battery, 2. a means of detecting the presence of electrical energy, such as the coil of a relay, 4-4

Overview of Railroad Signal Circuits

3. a medium through which to transmit the electrical energy from the source to the detector - in such a way that the energy will be noticeably changed by the presence or location of a train (i.e. the rails of the railroad track), 4. an output which provides information about the presence or location of a train to other parts of the railroad signaling system (e.g. the contacts of a relay). In railroad signaling, the medium that provides a means of transmitting the energy from the source to the detector consists primarily of the rails of the track, but also includes the dirt and rocks under the tracks, the wheels and axles of railroad locomotives and cars, rail joints, bondwires, gauge-rods, gauge-plates, railroad ties, tie plates, and many other seemingly mechanical components. Even though these parts are physically large, somehow they often get missed in a first introduction to signal systems. We will begin our explanation of track circuits by building up a track circuit, piece-by-piece, starting with its most basic mechanical components. Railroad Rails and Ties Mechanical and Electrical Properties

Figure 4-1 Picture of Railroad Tracks

4-5

Overview of Railroad Signal Circuits

Railroad Tracks Railroad tracks are such a common part of our modern world that we often take them for granted, without recognizing their individual components. If we stop to take a detailed look at railroad tracks, the first two things we notice are usually the rails and the ties. Next we may see the railroad spikes or screws attaching the two. However, there are many other track components that play important roles in railroad signaling, even though they may appear to have only a mechanical purpose at first glance. First, consider the steel used for the rails. Modern railroad rails are made from very tough Manganese-alloy steel. This alloy is slightly more than 97% Iron, and contains roughly 0.70% to 1.25% Manganese, 0.67% to 0.80% Carbon, and 0.1 to 0.5% Silicon. The rail steel may also contain small amounts of Nickel, Chromium, Molybdenum, or Vanadium. The undesired elements Phosphorus and Sulfur are held to maximums of 0.035% and 0.037%, respectively. Most of the railroad rails currently in use are of approximately this composition. However, this basic formula is always evolving, with railroads, researchers, and rail manufacturers continuously improving and modifying it, in order to reduce the life-cycle costs and improve the performance of the rails. Rail steel has a high tensile strength (roughly 120,000 p.s.i. to over 200,000 p.s.i. for some new rail steels1), considerable work-hardening characteristics, and is designed to handle the heavy impact and cyclic loading that railroad rails must withstand. Rail steel is usually rolled into rails having an industry-standard cross-section or profile, although anyone can have a special rail of their own design made for them by any of the major rail manufacturers - provided they are willing to pay for it. This has happened throughout the history of railroading, as railroads have searched for more durable rails to carry the everincreasing axle weights of their trains. But industry-standard profiles are by far the most common ones found on railroads in North America today. The standards for these profiles or sections, as they are called, were formerly maintained by AREA, the American Railway Engineering Association, which merged with several other railroad industry organizations on October 1st, 1997 to form AREMA, the American Railway Engineering and Maintenanceof-way Association. The standard AREMA rail sections are listed by weight, using the approximate weight (in pounds) of a three-foot (1 yard) length of rail made in accordance with the dimensions specified for that particular rail section. Standard sizes include 115 lb., 119 lb., 132 lb., 133 lb., 136 lb., and 140 lb., with 115, 133, and 136 being among the most commonly used today.

1,000 p.s.i. equals 1 k.s.i.

4-6

Overview of Railroad Signal Circuits

Figure 4-2 Typical AREMA Rail Section (133 lb. Rail)

The three major parts of a rail are the head, the foot, and the web. The head consists of the thick upper portion of the rail, which can withstand the intense pressures from the wheels of the train. This is also the part that experiences most of the wear when in service. The foot is the broad base of the rail, which distributes the weight of the train over a much larger area. Obviously, the web is the thinner portion of the rail in-between the head and the foot. It provides much of the flexural stiffness of the rail.

Figure 4-3 Thermite Weld in Rail

4-7

Overview of Railroad Signal Circuits

Steel rails can be welded via thermite welding, flash welding, or one of several arc-welding processes. All of these types of processes can be used in maintenance operations, with flash and thermite welding being the primary processes used in the construction of new track. Thermite welding is economically attractive, in that it can be performed in the field by a small crew (2-3 people) equipped with relatively inexpensive tools. Flash welding is much faster, and may provide a better weld, but requires the use of expensive equipment. Rails were once cut into 39-foot or 80-foot lengths at the steel mill, to allow for easy transport on a single railroad flat car. However, rails are increasingly being supplied in continuous lengths more than 500 feet long. These long ribbons of rail are carried on specially built sets of rail cars, with the pieces of rail spanning multiple cars. Using such extended lengths of rail minimizes the number of rail joints needed, and thus the number of welds that must be made in the field. Minimizing the number of rail joints reduces the maintenance costs, poor ride quality, and noise associated with jointed rail. Although most railroad tracks have only two rails, some may have as may as six. The two running rails (the ones the wheels roll on) are still spaced at the usual 56.5 inches apart (North American standard gauge), but other rails may be added to provide additional safety of one type or another. This is usually done where the tracks pass through a tunnel, over a bridge, or past some other potentially hazardous structure.

Figure 4-4 Function and Designation of Rails

Brace rails and check rails provide additional stiffness to the running rails, and are mechanically attached to them.

4-8

Overview of Railroad Signal Circuits

Figure 4-5 Guard Rails Minimize Damage Caused by Derailed Wheels

Guard rails are intended to help guide a derailed wheel which is rolling in-between the running rails. The running surfaces of the guard rail are not used, with the guard rail merely acting as a fence to limit the amount of lateral displacement of the wheel/axle. This helps avoid damage by keeping a derailed car of a moving train from straying too far from the centerline of the track. This is of particular concern in tunnels that curve, where partially- or fully-derailed portions of a train could shift towards the inside or outside of the curve, depending on the tensile or compressive forces present in that part of the train. A partially-derailed car could scrape along the walls of the tunnel, which could lead to a more serious accident. It is also not uncommon to see a pair of guard rails installed in tracks which pass under a highway bridge - extending for a hundred feet or so on either side of the bridge. In a worst-case scenario, these guard rails could prevent a partially-derailed railroad car from destroying the supports of the highway bridge, with potentially disastrous results. The standard rail gauge for North America is 56.5 inches. This is the nominal distance between the inside surfaces of the heads of the rails (known as the gauge faces). Knowing that engineers have a habit of choosing nice round numbers for standards, this number may seem a bit unusual - and it is. This gauge size may have deep historical roots reaching back to the time of the Romans - when chariots and other horse-drawn wagons were constructed with a common axle width. This practice made good sense at a time when muddy dirt roads were the norm, and the roads were often deeply rutted. The consistent axle width allowed the wheels of each cart or wagon to follow easily in the tracks of the one ahead of it. With little steering to be done, trains of wagons were assembled by connecting the reins of each animal or team to the rear of the wagon ahead of them. Eventually these trains acquired flanged steel wheels rolling on steel rails, and were pulled by a steam-driven locomotive. But the width (gauge) of the axles and tracks remained essentially unchanged.

4-9

Overview of Railroad Signal Circuits

Even though the standard North American rail gauge is 56.5 inches, railroad tracks are usually constructed a small fraction of an inch wider than this, in order to provide clearance between the gauge faces of the rails (the vertical sides of the rail heads which face the center of the track), and the flanges of the wheels on the locomotives and cars. This is done to reduce the amount of noise and friction generated by the flanges of the wheels against the gauge faces of the rails. The clearances and tolerances allowed for this vary, depending on the class of track that is being constructed or maintained. The rails are held at gauge by ties, spikes, tie plates, gauge rods, gauge plates, and other mechanical track components. The easiest of these to see are the large wooden timbers known as ties (the British call them sleepers). Although ties were once made almost exclusively of wood (oak in particular), ties made of steel-reinforced concrete, plastic, steel, and even exotic woods like Azobe have become increasingly common. Each material has its own particular advantages and disadvantages with respect to load-carrying capability, service life, cost, and other factors. Reinforced cast-concrete ties are most commonly used by commuter railroads carrying passengers due to their uniformity and durability, but are finding increasing use on heavy freight lines. Steel ties are most often used in specific locations such as bridges and tunnels, where their low-maintenance characteristics make them worth the additional cost.

Figure 4-6 Pictures of Railroad Ties, Wood and Concrete

The purpose of the ties is basically threefold. First, they must hold the rails at a fixed distance apart. Second, they must spread the weight of the train over a large enough area to prevent the track from being worked down into the earth with the passing of each axle. Thirdly, they must do both of the above without creating an electrical connection between the two steel rails (a shunt or short-circuit). This third requirement is what allows the track circuits to function properly. Wood is usually an acceptable insulator, at least at the voltages and currents used in railroad signaling circuits, but steel ties and steel-reinforced concrete ties must be specially constructed or otherwise insulated so that there is no electrical connection between the rails.

4-10

Overview of Railroad Signal Circuits

Figure 4-7 Bridges Pose Special Track Circuit Problems, Due to their (Conductive) Steel Construction

The rails do not rest directly on wooden ties, but instead rest on steel plates known as tie plates. These steel plates distribute the compressive load applied by the foot of the rail over a larger area of the tie, and provide pre-spaced holes through which to drive the spikes. They also help to tilt the rails slightly inward, and grip the foot of the rails to hold the rails at gauge.

Figure 4-8 Close-Up Pictures of a Tie-Plate, a Spike, and the Installed Assembly at the Foot of a Rail

4-11

Overview of Railroad Signal Circuits

Although they are an old symbol of the railroads, track spikes are still the most commonly used method of attaching rails to ties. Other methods use large screws or spring-steel clips to attach the rail to special tie plates, or special ties. In the case of concrete or steel ties, spring clips are used, along with a rubber or plastic insulator placed between the rails and anything else that might cause a direct electrical connection (a short or shunt) from one rail to the other. This can be of particular concern where the rails are attached to the beams of steel bridges. It should be kept in mind that steel or concrete ties can fail in a shorted manner if they are defective in their manufacture, thus placing an electrical shunt across the railroad track. This differs significantly from wooden ties which, although they can become more conductive when soaked or otherwise contaminated with conductive foreign materials, generally do not have the capability of effectively shunting the track.

Figure 4-9 Picture of a Pigtail Rail Clip and Insulating Pad on a Concrete Tie

The spikes, screws, or clips hold the rails and ties together, both vertically, and transverse to the track. However, the rails still have a limited ability to move along their length. That is, the rails can shift their position slightly along the route that the track follows. Curves and other structures tend to limit this somewhat, but some back-and-forth creep with thermal expansion and contraction is unavoidable. It usually amounts to less than an inch at any given point, as long as the overall track structure is intact. Normally, the effect of temperature on railroad rails is to change the amount of tension or compression within the rails themselves. In extreme cases, the longitudinal tensile stresses on the rails can exceed the tensile strength of the rail steel, and result in a broken rail or pull-apart. This usually occurs during the coldest winter temperatures.

4-12

Overview of Railroad Signal Circuits

If the rail is under compression, the strength of the spikes, ties, and tie plates securing the rails may be exceeded, causing the rail to break away from the ties and bow outwards, resulting in a sun kink. In some cases, the rail may even break when this happens. As would be expected, this kind of problem usually happens during very warm weather. To a mechanical or civil engineer, railroad track is a pair of flexible beams that serve to support and guide a train. The railroad ties and associated components serve to spread the weight of the train, in the form of pressure from the foot of the rail, over a larger area. The ties also serve to hold the rails in an upright position, and keep them separated by a uniform distance (the gauge of the track). And the ballast (the rocks underneath the ties) and sub-grade materials located around and below the ties can be viewed as elastic media with varying and sometimes unpredictable compressive and shear strengths. Uniformity is the key to safe and economical track. Any discontinuity in the tracks vertical section modulus (the amount of vertical deflection when under a load) will result in a soft or hard spot in the track. These discontinuities result in increased dynamic loading of the track structure, which diminishes the useful life of the rail, ties, and ballast. Higher train speeds and heavier axle loads greatly magnify the adverse effects of profile irregularities and poor track geometry. Ballast Supporting this varied assembly of rails, tie-plates, spikes, screws, clips, insulators, and ties is the most humble, and often the most overlooked, mechanical and electrical component of the track: the ballast. The term ballast can refer specifically to the crushed rock upon which the ties are laid. However, this term is also used generically to refer to the mix of rocks, dirt, sand, rain, snow, road salt, mud, grease, rust, organic solids, spilled loads, environmental contaminants, and assorted kinds of goo that make up the part of the earths surface that the tracks pass over. Even though railroad tracks are constructed with the intent of preventing an unintentional electrical connection between the rails, real-world ballast always provides some degree of electrical leakage from one rail to the other. Because of its mechanical properties in supporting the track, railroads use an enormous quantity of crushed rock ballast. Therefore they are often looking for the cheapest mechanically-suitable material they can get. Sometimes this leads to bad choices, like using crushed foundry slag containing an excess of conductive solids. Not all foundry slag is bad however, and some slag makes excellent ballast, especially when dry. It all depends on the electrical conductivity of the material used. Good ballast rock has a very low amount of electrical leakage or conductance, and therefore a high electrical resistivity or resistance. Being comprised primarily of metallic oxides and silicates, many types of rock easily satisfy this requirement. But the insulating properties of the ballast rock are always tempered by the presence of other more-conductive substances.

4-13

Overview of Railroad Signal Circuits

Figure 4-10 Modern Track Construction Techniques can Provide Very High Ballast Resistance. (U.S.&S Inc.) Here we see Rails Resting on Concrete Ties that are Themselves Set in Concrete, Instead of Crushed Rock

The resistance of railroad track ballast can be measured by a variety of similar techniques. But in each case the concept remains the same, as do the units of ballast resistance used by railroads. The most-widely used unit of measure for ballast resistance is Ohms per 1000 feet of track2. However, the correct dimensions for this are actually OhmKilofeet, or OhmKft, for short. The meaning of ballast resistance is best illustrated by an example. If we were to take a 1000-foot section of track with no other equipment connected to it, and isolate it from the rest of the railroad at both ends, and then measured the resistance between the rails with the proper type of ohmmeter3, the ohmmeter would indicate the ballast resistance in Ohms for 1000 of track. This is the ballast resistivity in OhmKft. To make this same measurement on an isolated section of track of any arbitrary length, we would simply multiply the reading on our special ohmmeter, and multiply the reading in Ohms by the length of the section of track as measured in Kft. In track circuit simulations, it is very important to distinguish between ballast resistivity, and ballast resistance. As either number may be used by personnel doing studies of the impact of a proposed ac power line on a nearby railroad line, attention must always be paid to the units (dimensionality in use).
2

The actual dimensions of track ballast resistivity are OhmKft, even though you will most often hear this described as Ohms per Kft or Ohms per 1000 feet on most railroads. A 1000-foot section of track built on ballast with a resistance of 4 Ohms per 1000 would have a resistance of 4 Ohms from rail to rail. A 2000-foot section of track built on identical ballast would have a rail-to-rail resistance of only 2 Ohms. Thus the expression Ohms/1000 feet is incorrect, as it would result in increasing rail-to-rail ballast resistances with increasing track length, when in fact the exact opposite is true. This misuse of units is simply a result of the oversimplification of ballast resistance to make it more understandable to new signal maintenance personnel (see example above). Due to electrochemical polarization effects, like those seen in the testing of ground rods, an ohmmeter using an ac signal source must be used for accurate measurements.

4-14

Overview of Railroad Signal Circuits

In the real world, ballast resistivity can vary greatly, from less than 0.25 OhmKft of track, to over 100 OhmKft. In general, most railroad track ballast is maintained at 2 Ohms/1000 or higher, in order to allow for the use of track circuits of reasonable length. Some railroads, with a thorough understanding of the effects of ballast resistivity, maintain their tracks to even higher standards of resistivity. This can help to avoid the nuisance activation of Crossing Predictors, which can become more susceptible to the effects of ac interference from power lines under conditions of low ballast resistivity. A ballast resistance of 4.0 OhmKft or higher will generally improve the operation and reliability of most track circuits, and is highly preferable. Automated processes for maintaining the ballast have helped improve the uniformity of the ballast, and have increased the minimum ballast resistance. Where used, concrete ties have also generally raised ballast resistances, and have been beneficial to track circuit performance. It should be recognized that ballast resistance is an average measurement, and there will always be places along a track circuit that will be significantly above or below this average. The most common place for an isolated section of low ballast resistance to occur is at grade crossings. This is due to the restricted drainage present at most grade crossings. The salting of roads in the winter also plays a significant role in contributing conductive materials to the ballast at a grade crossing. The poorer (lower) the ballast resistance, the shorter most track circuits must be, due to the shunting action of the ballast, which shunts some of the current applied to the track by the track circuits transmitter. Ballast resistance (leakage conductance) accounts for most of the loss in signal strength when transmitting an electrical signal via railroad tracks, particularly at lower railroad signaling frequencies from zero to a few hundreds of Hertz (Hz). The attenuation of most ac voltages transmitted through railroad tracks (and their associated ballast) can be very high, due to the added effect of the inductive impedance of the tracks. However, ballast resistance still accounts for the majority of the loss. Some very long ac track circuits operate with as much as 70 dB of attenuation between the transmit and receive ends of the circuit. Contaminants in Ballast Although ballast materials themselves are chosen for their mechanical and electrical properties, these can both be affected by the presence of other materials, which can become mixed with the ballast over time. These may include: Road Salt (Particularly at Grade Crossings) Sand and other inorganic fines Spilled bulk chemical loads such as coal, industrial chemicals, or fertilizers Spilled grain, processed agricultural products such as soybean meal, or cattle feed

Anything that is inherently conductive, or that can form a conductive solution in water, or can help the ballast rock retain water, can affect the electrical characteristics of the ballast. Many of these ballast contaminants are long lasting, and have deep historical roots.

4-15

Overview of Railroad Signal Circuits

The substitution of mechanical refrigerator cars for ice-cooled reefers in the early to middle 1960s eliminated the longstanding plague of salt-brine drippings in switching yards and rail terminals. This conductive solution can produce extremely low ballast resistances. And the retirement of large stoker-fired steam locomotives has reduced the presence of other conductive materials in the ballast, such as carbon and ash, which were sometimes discarded along the tracks. But blackish deposits in the ground, thought to be locomotive firebox coals and ashes, can still be found just below the surface at locations along older railroad lines. Railroad Track as an Electrical Transmission Line In this section, we will begin discussing the characteristics of railroad track, not as a mechanical component of a transportation system, but as an electrical conductor. Specifically, we will look at railroad track as an electrical transmission line. This is not a transmission line in the sense of a high-voltage line used to transmit power, but a transmission line in the sense used by radio and communications engineers. The electrical characteristics of railroad track are what make all of the different types of track circuits used for railroad signaling possible. There are basically two types of rail joints: insulated rail joints (called insulated joints or just I.J.s for short), and un-insulated or ordinary rail joints (simply called rail joints). Both types of joints mechanically connect the ends of two pieces of rail together, but insulated joints do so while electrically insulating the two rails from one another.

Figure 4-11 Ordinary and Insulated Rail Joints

From the earliest days of railroading, the familiar sound of a passing train was produced primarily by the rhythmic clickety-clack of the trains wheels rolling over the small gaps between the ends of the individual pieces of rail at each rail joint. These gaps were both necessary and unavoidable, as they provided room for movement as each piece of rail expanded and contracted with temperature. Keeping all of these bolted rail joints in good condition despite the passage of trains and temperature fluctuations was a significant track maintenance chore. Although the ever-increasing use of continuous welded rail has greatly reduced the number of rail joints used on railroad tracks, it has not completely eliminated them. 4-16

Overview of Railroad Signal Circuits

Bolted rail joints provide a mechanical connection, as well as some degree of electrical connection (generally poor and variable), from one piece of rail to the next. The joints are constructed using a strong steel joint bar on either side of the rails to be joined, with the joint bars bolted to each rail end with two or more large bolts on each side of the joint. These bolts pass through holes drilled or punched in the joint bars, as well as the holes drilled in the webs of the rails to be joined. Bondwires In order to work reliably, signal circuits that use the rails as electrical conductors require that the rails within each circuit be connected together electrically in a low-resistance and reliable fashion. Because the impedance of the electrical connection at an ordinary rail joint is much higher than optimum for track circuits, and can vary widely as the pieces of the joint shift their positions relative to one another over time under the influences of temperature and passing trains, a consistent low-impedance flexible connection known as a bond wire is attached between the ends of the rails joined by a rail joint.

Figure 4-12 Welded Bondwires and Mechanically-Attached Bondwires

Sometimes called chickenheads or chicken-necks, mechanical rail connectors come in several different varieties, most of which have a tapered portion which is driven into a cylindrical hole drilled through the web or into the head of the rail. They also have a means of connecting a wire (roughly #6 AWG or larger) to the pin using a crimped connection of some type. These can also used to attach the wires of the track circuits to the rail. Another way of attaching a bondwire or rail connection is the welded connection. Often referred to generically as Cadweld (Cadweld is a registered trademark of Erico International Corporation), this form of bondwire connection welds a short pigtail wire directly to the head or web of the rail using aluminothermy and a clamped-on mold. As a permanent connection to the rail, welded connections for bondwires and track wires are often more reliable than mechanical rail connectors. Some signal equipment manufacturers strongly recommend their use with their signal equipment. 4-17

Overview of Railroad Signal Circuits

Any type or style of electrical connection to the rails would be suitable, as long as it provides a consistent low-resistance connection from the wire to the rail. In the case of bondwires, which connect the adjacent ends of two rails together at a rail joint, there may literally be hundreds of welded connections within a track circuit. Over time, one or more of them will eventually fail, given enough years of service. When they do fail, it is usually due to gross mechanical damage or corrosion. But some simply work themselves loose from the rail due to improper installation and the normal flexing of the rails as each axle of the train passes by. When the behavior of a track circuit constructed using jointed rail and bondwires becomes erratic or unreliable, experienced signal maintainers know to check their bondwires first. This is sometimes called walking the bonds. This means performing a physical and visual examination of every single bondwire within the affected track circuit. By pulling on each bondwire with an improvised tool, constructed of a length of locally-available recyclable material (commonly known as a stick), the maintainer tests the mechanical integrity of each connection. All too often, the cause is found to be a broken bondwire damaged by vandals. Fortunately these are easy to find and fix. More difficult is the case of a corroded bondwire that has lost most of its strands to corrosion inside of one of its crimped or welded ends. The high resistance created by a corroded or otherwise damaged bondwire can be very difficult to find. This is why the inspection is both visual and mechanical. In the early days of railroading, bondwires were made of iron, and were mechanically attached to the web of the rail with a channel pin. These pins held the bond wire in a groove in their side, which collapsed and held the wire tightly when the pin was driven into a hole in the web of the rail. The limited conductivity of the iron, as well as its propensity to rust, frequently contributed to high resistance connections. High-resistance bondwires can create many different problems, some of which can mimic the symptoms of ac interference. Bad bondwires can also lead to a loss of broken rail detection capability in the vicinity of rail joints, as well as a host of other problems. Todays welded rails and welded copper bond wires can almost eliminate the effects of bond wire resistance from track circuit calculations and considerations. Bond wires still come in several styles, including long bonds, railhead bonds (both welded and pin-type), and highcurrent bonds. Some of them are mechanically attached to the rails and coated with a protective layer of anti-oxidation grease, and some are welded to the rails. Examples of several of these are shown below.

4-18

Overview of Railroad Signal Circuits

Figure 4-13 Jointed Rails with Railhead Cadweld Bonds, Long Bonds, and High-Current Bonds

4-19

Overview of Railroad Signal Circuits

Rails Regardless of what size rails are used, or in what form they have been supplied, or how they have been electrically bonded together to form continuous conductors, the railroad rails themselves are generally good conductors of electricity. This is unsurprising, given their large cross-section. For example, the cross-sectional area of an AREMA standard 133 lb. rail is just over 13 square inches. This equates to about 16.6 million circular mils. Even though the -5 4 electrical resistivity of the rail steel is approximately 7.0 x 10 Ohm-centimeters , or roughly -6 41 times the resistivity of copper (1.72 x 10 Ohm-cm), the massive cross-sectional area of each steel rail gives it a resistance roughly equal to two #4/0 AWG (pronounced: four-ought) conductors in parallel. This works out to roughly 0.03 Ohms per 1000 of rail. So although high-manganese rail steel might not be an electrical engineers first choice of materials for an electrical conductor, it can still be a good one if the conductor is large enough in cross-sectional 5, 6 area . The Shunting Action of a Trains Axles The idea that the solid steel axles and wheels of the train provide nearly a dead short from one rail to the other is presently the basis of almost all train detection and railroad signaling in North America and parts of the U.K. and Australia. In the United States, the FRA even specifies the required minimum sensitivity of all track circuits used in rules 234.229 and 236.56, requiring the track circuits to be able to detect the presence of a shunt placed across the track, of 0.06 Ohms. This is a maximum resistance value, with track circuits also detecting all shunts that are lower in resistance than this.

It should be noted that the values given in the literature for the resistivity of rail steel vary widely. When combined with the various rail sizes (weight per yard) used, an even wider range of resistances per unit length of rail is created. For most calculations, the resistance of the rails has an insignificant effect on the results, but a value of 0.03 Ohms per 1000 of (single) rail is often used. Other values may differ from this by a few orders of magnitude. Even though the dc resistance of railroad rail is quite low, it still exhibits significant impedance to an alternating current. Some have approximated this impedance as roughly one Ohm per Kilometer of rail at 60 Hz, but this figure varies somewhat in the available literature. It is for this reason that ac track circuits often use the lower audio frequencies (below 1 KHz). Electrically speaking, railroad track can be viewed as a lossy but workable communication transmission line, not unlike the old 300-Ohm twin-lead antenna wire once used to connect antennas to television sets. However, due to the large differences in size and other physical factors, standard gauge railroad tracks have a calculated characteristic impedance that is roughly four times as high (1200 Ohms). However, for the railroad EMC specialist, the characteristic impedance of the track is rarely of concern for most track circuits. What is of concern for track circuits is how the electrical characteristics of the transmission line (the railroad track) change when a short or shunt is placed across it. This is the characteristic measured by the most commonly affected railroad signal circuits, i.e. Motion Sensors and Crossing Predictors. Many Electrical Engineers have wrongly assumed that Motion Sensors and Crossing Predictors operate on the principle of Electrical Time-Domain Reflectometry, or some other sophisticated method. They actually just measure the magnitude and phase of the impedance of a length of track with a shunt from rail-to-rail at a point some distance away from the point of measurement.

4-20

Overview of Railroad Signal Circuits

Many track circuits perform train detection by transmitting an electrical signal from one end of an independent section of track, and monitoring the strength of the received signal at the opposite end. This electrical signal is often essentially constant in power, voltage, or current level7.

Figure 4-14 Unoccupied Track

Placing a shunt across the track anywhere between the transmitter and receiver effectively shorts out the signal from the transmitter, and prevents it from reaching the receiver. Electrical Engineers call this reduction in signal level insertion loss. When the insertion loss of the shunt drops the signal level at the receiver below a certain threshold, the track circuit then declares that section or block of track to be occupied.

Alternating or otherwise time-varying currents are used in all track circuits except simple dc track circuits. Furthermore, the carrier may be modulated using any combination of amplitude, frequency, or phase modulation. This modulation is often necessary in order to make sure that the receiver will only respond to electrical signals from the correct transmitter.

4-21

Overview of Railroad Signal Circuits

Figure 4-15 Occupied Track

This is indicated electrically by the position of the track relays armature (see figures above). Most track circuits operate with a rail-to-rail voltage of 3 volts or less, using either alternating or direct current at as much as a few amps, but usually much less. The most notable exceptions to this generalization are the high-energy track circuits used where there are insulating or semiconductive films (rust, spilled loads, etc.) present on the running surfaces of the rails. These films can prevent the wheels of the trains from making adequate electrical contact with the rails. The higher rail-to-rail voltages used by these track circuits can effectively punch through the film on the rail, and improve the ability of each axle to provide an effective shunt to the track circuit. Although dc track circuits were developed first, since that time there have been a number of different types of track circuits that use ac electrical signals to detect the presence of a train. Many of them are used for simple presence detection, just as dc track circuits are. These track circuits normally use an ac signal that usually falls within the audio frequency range of 20 Hz to 20 KHz, although frequencies outside of this range have also been used. The limited bandwidth of railroad track tends to discourage the use of higher frequencies, and generally creates an inverse relationship between the length of the track circuit, and the highest useable operating frequency for the ac track circuit. Simply stated, higher frequencies are generally used for shorter track circuits, and lower frequencies are generally used for longer track circuits. Other ac track circuits such as Motion Sensors and Predictors view the track somewhat differently. These are devices which detect trains at highway grade crossings, where railroad tracks cross a highway, road, street, or even a pedestrian walkway.

4-22

Overview of Railroad Signal Circuits

Figure 4-16 Grade Crossing

They have transmitters and receivers that are located very close together, sometimes being connected to the rails within a few inches of each other. These systems view the track as a shorted transmission line, whose physical length (distance from signal equipment to the short or shunt) can be measured. When a transmission line made from two low-resistance parallel conductors is shorted (shunted) at one end, the capacitance of the line effectively disappears, and all that remains is the inductance of the conductive loop that has been formed. In the case of railroad tracks, our loop is 56-1/2 inches wide, and as long as the distance from the receiver/transmitter at the crossing, to the shunt. The magnitude of the impedance (primarily composed of inductive reactance) of this long narrow loop of rail is roughly 3 Ohms per KHz per 1000 of track, depending slightly on the frequency used and other factors such as ballast resistance. Shunted railroad track loops have been observed to have between 0.43 and 1.57 microhenries of inductance per foot of track (varies with frequency, etc.), with 0.47 microhenries being a good value to use for most calculations involving Motion Sensors and Crossing Predictors which are operating at frequencies between 86 Hz and 970 Hz. Since track circuit engineers deal with thousands of feet of track in each signal circuit, this is often thought of as 0.47 millihenries per 1000 feet of track. The exact value of this inductance varies with the frequency, the weight (section) of rail, the magnetic permeability of the rail steel, the physical composition of the ballast and ties, and other factors. Motion Sensors and Predictors measure the impedance of the loop formed by the two rails of the track and the axle(s) of the approaching train by applying a constant-current, audio-frequency ac signal to the track, and performing several calculations based on the ac voltage which results. Since the magnitude of the resulting ac voltage is proportional to the impedance of the loop, and 4-23

Overview of Railroad Signal Circuits

the impedance is proportional to the distance from the receiver/transmitter to the train, the magnitude of this voltage can be used to determine the trains position with respect to the Motion Sensor or Crossing Predictor installed at the grade crossing. By measuring the way that this voltage changes over time, the velocity of the train can also be calculated.

Idealized Impedance vs. Train Position (unidirectional approach)


100 (% of unoccupied impedance)

80

60

40

20

0 100 80 60 40 20 0 Train Position (% of approach unoccupied)

Figure 4-17 Impedance vs. Train Position

In order to establish the outer limit(s) of the area within which the Motion Sensor or Predictor will look for approaching trains, a termination shunt is connected across the track (from one rail to the other) at a location some distance away from the crossing. This also places an upper limit on the impedance of the track circuit used by the Motion Sensor or Predictor. Termination shunts are usually made from a series-resonant Inductor-Capacitor (L-C) combination, but may simply consist of a very large capacitor (>100,000 uF), or a short length of heavy-gauge wire ( #6 AWG). Thus, we can see that railroad track can be used as a transmission line for carrying electrical signals from a transmitter to a distant receiver (simple occupancy determination), or as a variable-inductance loop for measuring a trains exact location and/or velocity relative to a grade crossing (crossing prediction or motion sensing). In the real world, it is often used in more than one way at the same time. Probably the most important aspect of railroad track circuits is that the track itself is used as part of the circuit. In this way, a single circuit can be used not only to detect the presence of a train, but also to simultaneously check the physical integrity of the track itself.

4-24

Overview of Railroad Signal Circuits

For example, if a rail were to crack and pull apart in cold weather, the amount of current arriving at the receiver would drop to nearly zero, and the track circuit would declare the block of track to be occupied. And the same thing would happen if an unintended short (false shunt) developed across the track between the transmitter and receiver. The received signal level would drop to nearly zero, and the block would be declared occupied, turning the signal lights at each end of the block to red. As trains normally are not permitted to enter an occupied block, as indicated by the red signals, the occurrence of either a short or an open in the track makes the system fail safe. Insulated Joints Although there is a maximum physical length for each different type of track circuit, which is determined by its design and the laws of physics and safety, there are many reasons for constructing track circuits of much shorter length. As most track circuits simply determine whether or not that section or block of track is occupied, the ability of a signal system (or train dispatcher) to determine the exact location of a train is limited by the number of track circuits used. For example, we could imagine a railroad constructed so that the entire railroad was contained within the boundaries of a single track circuit. This would probably require a very small railroad, a very large track circuit, or perhaps both. Even with just a single locomotive or railroad car on the railroad, the track circuit would always appear occupied. This sort of information isnt very useful to a train dispatcher or anyone else. But if we split the railroad into two separate track circuits, each of which were large enough to completely contain the locomotive or railroad car, say, a bit more than 100 feet long, then the track circuits could tell us which half of the railroad the locomotive was currently occupying. If we scale this concept up to the size of even a small short-line railroad, the number of track circuits needed for adequate resolution of a trains position quickly grows into the dozens. And a large railroad will have thousands of track circuits along its routes. Although there are track circuits that can operate at different ac frequencies to avoid interfering with one another, and can be connected to the same tracks simultaneously, sooner or later we will run out of different frequencies to use. The only solution to this problem at present is to partition the tracks into electrically separate pieces, and get a fresh start in each separate section. In order to electrically divide the tracks of the North American railroad network into pieces of a manageable size, insulated rail joints are used. These joints are similar to the usual rail joints, which use joint bars and bolts to mechanically connect the ends of two rails. However, insulated rail joints differ in one important respect. While they provide a solid mechanical connection between the two rails being held end-to-end, they do not provide an electrical connection. Insulating sleeves known as thimbles are placed over the shafts of the bolts, and a roughly 3/8" thick block of insulating material (usually phenolic) is placed between the ends of the rails (called an end post). Insulating materials are also applied between the joint bars and the rails themselves. Often adhesives are used to complete the assembly, and unite it into a single mechanical unit known as a bonded joint.

4-25

Overview of Railroad Signal Circuits

Some types of insulated joints can be assembled in the field from their individual components. However, insulated joints are increasingly being supplied in the form of a prefabricated insulated joint installed between two short (< 10 foot) lengths of rail. Rather than drill and prepare two rail ends in the field, a length of rail (< 20 feet long) is sawn out of the existing track with an abrasive cutoff saw and removed. The insulated joint assembly is then installed in its place, and welded. Insulated joints are usually installed in pairs, with one insulated joint in each rail of the track, normally within a few feet of one another. In nearly all instances, insulated joints are installed only where necessary, as a component of the track circuits used for block signaling, interlockings, or grade crossing warning systems. Insulated joints, especially failed ones, can play a key role in EMC problems on railroads.

Figure 4-18 Typical Insulated Joint Installed in Rail, with Polyurethane-Encapsulated Joint Bars

The design of an insulated joint represents a necessary engineering compromise between mechanical strength and insulating properties. Its presence in the track structure creates a structural discontinuity that tends to impair both the strength and long-term physical stability of the track. For this reason and others, railroads seek to minimize or eliminate the use of insulated joints whenever possible. Each insulated joint installed creates an additional maintenance expense, which adds significantly to the joints initial cost of purchase and installation. Although insulated joints are an important electrical component of many track circuits, they are still considered primarily as mechanical components of the track by most railroads. As such, they are usually the responsibility of the railroads track department. This means that when a signal department employee finds one or more failed (shorted) insulated joints, they must receive the cooperation of the track department in order to get the joint(s) replaced. Arranging for the replacement of an insulated joint can require additional time. This can create a significant delay in the correction of an ac interference problem caused by a failed insulated joint. 4-26

Overview of Railroad Signal Circuits

The design of most track circuits allows them to withstand the failure of one or more of the insulated joints isolating the track circuit from the rest of the railroad trackage. This has sometimes led to a stated or un-stated railroad policy of not dispatching track maintenance personnel to replace an insulated joint until the operation of the signal equipment has clearly been affected. When enough insulated joints have failed, to a degree sufficient to affect the track circuit, and there is a clear indication of a problem (e.g. one or more track circuits are indicating that the track is occupied, even when it isnt), the track department will be dispatched to work with the signal department to find and correct the problem. The problem with this approach to Insulated Joint maintenance is that it only takes a single shorted insulated joint to electrically imbalance the track. If there is a strong electromagnetic field at the rails, caused by nearby ac power lines, this imbalance can result in unusual rail-toground and rail-to-rail voltages on the track, which can interfere with the operation of the railroad signal equipment. In a worst-case maintenance scenario, the signal and track departments will work together just long enough to find and replace one of the two failed insulated joints. This is usually enough to return the signal system to normal operation, but leaves the signal system only a single insulated joint failure away from its next case of trouble. Insulated joints are simple in concept, but quite diverse in their reasons for use. Although they are always used to provide a mechanical end-to-end connection of rails while providing electrical isolation, the various reasons for their application include: Electrical separation of adjoining track circuits; Insulation of intersecting rails (as at a switch or diamond) to prevent the short-circuiting of a track circuit; Electrically transposing the rails in order to maintain the desired polarity at the opposite end of the track circuit; Mitigation of induced current in the rail, i.e. dividing the track into physically shorter signal blocks (reduces the induced ac voltage); Dividing or isolating electric propulsion current return paths; Balancing unequal rail resistances and currents (used with impedance bonds in electric propulsion territory).

The strength and quality of Insulated Joints have improved since the widespread deployment of Continuous Welded Rail (C.W.R.), which began in the late 1960s. The older styles of conventional Insulated Joints, lacked sufficient strength to resist the enormous longitudinal forces that can develop in continuous welded rails during rail temperature extremes. The most successful solution to this problem to date has been the bonded insulated joint. The bonded insulated joint is secured with both epoxy adhesive and bolts, in order to provide a high shear strength between the rails and joint bars. The mating surfaces of the rails and joint bars are carefully prepared by sandblasting to ensure high bond strength. The joint bars are electrically insulated from the rails by a fiberglass mesh that is impregnated with adhesive. The bolts are carefully tightened to specified torques to avoid damage to this insulation.

4-27

Overview of Railroad Signal Circuits

Figure 4-19 Cross-Section View of a Typical Bonded Insulated Joint (Portec Rail Products, Inc.)

The bonded joint assembly is delivered to the field as a pre-assembled unit, with short pieces of rail on either side of the joint. The bonded insulated joint is installed by cutting out a length of rail equal to the length of the bonded insulated joint assembly, and replacing it with the assembly. The rails are then welded or (rarely) bolted to the bonded insulated joint assembly. Because the bonded Insulated Joint cannot be repaired in the field, its higher cost limits its use to heavily used mainline trackage where the additional reliability is worth the additional cost. Non-bonded insulated joints continue to be widely applied. Fiberglass and polyurethaneencapsulated insulated joints are the most common types used for new installations or renewal work.

Figure 4-20 Sectional View of Rail and Typical Encapsulated Insulated Joint (Portec Rail Products, Inc.)

Other types of insulated joints such as the laminated densified-wood insulated joints (e.g. the Permali joint) have been widely used for many years, and can sometimes still be found on older rail lines. Due to the still-excessive failure rate of most modern insulated joints in heavy freight territory, manufacturers have continued to develop even stronger and more durable insulated joints.

4-28

Overview of Railroad Signal Circuits

Insulated Joint Failure Modes As is the case with other components of the signal system, insulated joints may fail mechanically, electrically, or both. Railroad signal systems must be designed to ensure that 8 electrically failed Insulated Joints do not create the risk of a false clear condition. Perhaps because of this design characteristic of railroad track circuits, the FRA rules do not require periodic testing of insulated joints. Instead, they require only that the insulated joints be maintained so as to . . . prevent current from flowing between the rails separated by the insulation in an amount sufficient to cause a failure . . . of the track circuits (rules 234.235, and 236.59). The only requirement above and beyond this is found in the section of the FRA rule book devoted to Grade Crossing Signal System Safety, where rule 234.271 specifies that Insulated rail joints, bond wires, and track connections shall be inspected at least once every three months. Surprisingly, no additional clarification of what constitutes an acceptable inspection is provided within sections 234, 235, or 236 of the rulebook (revised July 8, 2005). Under ideal conditions, the electrical resistance of an insulated joint should be nearly infinite at ordinary signal system voltages. This makes the testing of Insulated Joints, prior to their installation, relatively simple and foolproof. However, the accurate measurement of Insulated Joint resistances, once installed in the track, is much more difficult. This is due to the presence of electrical leakage paths through the ballast and around the insulation of the Insulated Joint. Although there are commercially available Insulated Joint Testers which apply an ac voltage of a frequency in the range of 3 to 60 KHz across the insulated joint, and measure the resulting current (insulation resistance measurement), these are still subject to inaccuracies caused by the leakage paths mentioned above, and require careful interpretation of the data they provide. The reason for using such high test frequencies is twofold: The poor high-frequency transmission characteristics of the rail help to minimize the influence of ballast leakage paths on the measurement. These frequencies are above the range normally used by railroad signal equipment, and generally will not interfere with the operation of the signal system.

There are also other techniques for testing insulated joints installed in railroad tracks - some of which involve the use of clamp-on ammeters installed directly around the insulated joint. The advantage of this type of technique is that it only measures the current flowing through the insulated joint, and ignores the current following a leakage path through the surrounding ballast. However, instrumentation of this type is not commonly found on railroads.

A false clear is a condition in which a train is given a signal indication which is more permissive than it should be. This includes being allowed to move when it is not safe to do so (green or yellow vs. red), as well as being allowed to go faster than is actually safe (green vs. yellow).

4-29

Overview of Railroad Signal Circuits

When insulated joints fail in the electrical sense, it only means that the insulation they provide is now less than adequate (there is no upper limit on the insulation resistance of a good insulated joint). The most common causes of insulated joint failure are: Accumulation of metallic dust or slivers across rail ends. The battering effect of railroad wheels rolling across the gap between the rails tends to magnetize the ends of the rails. This in turn tends to attract ferromagnetic particles. Burrs and sharp edges from rail cutting or bolt-hole drilling. This condition is often difficult to diagnose, as it can be hidden behind the joint bars. Loose, missing, or improperly installed components. Joint run-together due to expansion and movement of rails. This can happen when the insulating end-post between the rails is old, missing, or severely compressed. Improper installation. Sometimes a track crew will mistakenly install an insulated joint so that the gap between the rails rests on top of a metal tie-plate, thus bridging the gap. In other cases, rail spikes have been installed so as to bridge the gap. Permanent dielectric breakdown due to excessive voltages from lightning, etc. Rail Batter. Over time, the battering effect of the wheels of passing trains can actually deform the heads of the rails enough to make them touch. This is always a result of poor track maintenance.

Figure 4-21 Metallic Dust and Slivers Tend to Accumulate, Bridging Rails at Insulated Joint

4-30

Overview of Railroad Signal Circuits

Track Circuits Except for a few experimental and alternative techniques for detecting trains, almost all train detection in North America is done via track circuits. This contrasts sharply with European signal practices, which rely much more heavily on so-called axle-counter techniques (discussed below). A variety of track circuits can be used to provide a fail-safe means of determining the presence, location, length, and velocity of trains. In order to better understand the details of how track circuits work, we will begin with the basics. At its simplest, a track circuit is simply a low-voltage dc electrical circuit, which uses railroad mechanical components such as rails, wheels, and axles as some of its electrical conductors. These components are used in conjunction with a relay and a battery, connected to the tracks, in order to detect the presence of railroad axles, cars, or locomotives within the track circuit. This track occupancy information is then used by the signaling systems at wayside signal locations, interlockings, and some crossings to control the operation of switches, wayside signals, flashing red lights, bells, gate arms, and other devices and systems. What Is (and What Isnt) a Track Circuit? The FRA defines the track circuit as: A circuit of which the rails of the track form a part. Unfortunately, this definition isnt narrow enough to be useful. For example, the trains used in many so-called light rail commuter systems are propelled by electric motors which obtain power from either an overhead catenary wire, or an energized third rail. In each case, the current supplied to the traction motors must have a source path, and a return path. The source is obvious, but what many people fail to realize is that the tracks themselves are used as the primary return path. Some of the current may return through the earth to the transformer at the propulsion substation(s) supplying the power to the train, but most of it returns via the rails. Yet propulsion current circuits are not considered track circuits. Rails often form a part of grounding circuits intended to mitigate or equalize voltage potentials between the train, track, and surrounding structures, especially where electric traction systems are in use. But these circuits are not track circuits either. Another example of a circuit that is connected to the rails, but is still not a track circuit, involves the electric resistance heaters (called snow melters, or switch heaters) installed at track switches. These are used to keep the movable parts of the track at track switches from either freezing together, or not moving correctly due to blockage by snow or ice. The heating elements of these heaters are often deliberately grounded to the rails of the track, but these are still not track circuits. Where railroad signaling systems are concerned, a track circuit is an electrical circuit that performs one or more specific functions such as: Detection of trains, cars, or axles Detection of broken rails or false shunts Communication of signal control information between wayside locations 4-31

Overview of Railroad Signal Circuits

Route locking of switches, moveable bridges, etc. Communication of cab signal or ATC information from wayside-to-train Communication of non-vital train ID data from train-to-wayside Detection of a trains presence and velocity at grade crossings Distance-to-couple measurements in switching yards.

Robinsons Patent Todays railway signaling technology could not have achieved its high degree of safety and reliability without one key invention: the track circuit. On August 20, 1872, a United States patent (see figure below), number 130,661 was issued to William Robinson for an electrical track circuit.

Figure 4-22 Robinsons 1872 Track Circuit Patent

4-32

Overview of Railroad Signal Circuits

As with many inventors of his time, Robinsons work was based on trial-and-error, aided by good luck, rather than studious and organized scientific inquiry. Robinson sold his interest in the track circuit to George Westinghouse in 1881, and did not make any further significant contributions to the field of railroad signaling prior to his death in 1921. Although track circuits were widely accepted, technical advancement was slow at first, with detailed mathematical analysis of track circuit parameters beginning in earnest in the 1920s. Robinsons track circuit consisted of a battery, a relay, and a length of track that was isolated from the rest of the railroad by the use of insulated rail joints. The terminals of a primary battery were connected, one to each rail, at one end of the circuit. At the opposite end, the coil of a relay (track relay) was connected across the rails and received energy from the battery through both rails, while the track was unoccupied. Although Robinsons method of using this circuit to control a wayside signal may differ slightly from modern techniques, this was still the humble beginning of all modern track circuits. We will refer to the normal operation of a track circuit in terms of the Transmitter, the Track 9 (both rails) or Rail (singular), and the Receiver . A basic dc track circuit consists of a Transmitter (the battery or other power source), and a Receiver (the relay), connected together via the rails of the track, as shown below.

Figure 4-23 Basic DC Track Circuit

When the current flows in an uninterrupted fashion from the battery, through the rails, to the coil of the relay (i.e. when there is no train present within the block), then the coil is energized, causing the relays armature to be attracted toward the relay coil. This action causes the relays armature and moveable contacts to make contact with one of two sets of fixed electrical contacts.
9

Conceptually, such a track circuit could be viewed as fitting a Source-Path-Receptor model, with the battery as the source, the rails as the path, and the relay as the receptor. Thus, the presence of a train is detected by the observed change in path characteristics when the train occupies the rails of the circuit.

4-33

Overview of Railroad Signal Circuits

The electrical circuits opened and closed by these relay contacts can then be used to control other signal equipment, such as wayside signals, cab signals, or grade crossing warning systems. They can also be used as inputs to the vital logic of a complete signal system.

Figure 4-24 Basic Track Circuit Schematic

Most often, we think of track circuits as being what controls the aspect of the wayside signals. Wayside signals are normally located along the side of the track, and are used to display a particular aspect which conveys information, called a signal indication, to the crew of a train as to what train movements may safely be made. Wayside signaling will be discussed in much greater detail, below. The important thing to remember is that it all starts with the track circuit. Here is an example of basic track circuit operation. If no train is present within the track circuit, a green signal light at the entrance to the block could be lit via the heel and front contacts of the track relay10. This would tell the crew of any train approaching that track circuit that there were no trains, cars, or axles occupying the track circuit. When at least one axle of a train, locomotive, or railroad car enters the track circuit, it will shunt (most of) the current coming from the battery, and will prevent it from reaching the coil of the relay at the other end of the track circuit. With almost no current energizing the coil of the relay, the armature of the relay will move to its de-energized position. The contacts closed in this position could then be used to light up a red signal light at the end of the block, telling other trains not to enter, as the track circuit was already occupied.

10

The terms front and back are used to refer to the contacts of the relay which are touched by the armature contacts in the energized and de-energized positions, respectively. The armature contacts are referred to as the heel contacts.

4-34

Overview of Railroad Signal Circuits

Figure 4-25 Occupied DC Track Circuit

One assumption, which underlies this technique of train detection, is that it only takes one axle to shunt enough current away from the coil of the relay, and move the relays armature from its energized position, to its de-energized position. Since the wheels and axles of most railroad cars and locomotives are assembled by pressing two steel wheels onto a solid steel shaft using hundreds of tons of pressure, the finished assembly actually does have a very low resistance or impedance from the tread11 of one wheel to the tread of the other. This can and does form an effective low-impedance path from one rail to the other, provided that the tread of each wheel makes good electrical contact with the rail. In practice, track circuits are adjusted so that the presence of a shunt of 0.06 Ohms or less, anywhere within the physical boundaries of the track circuit, will always result in the coil of the track relay being de-energized, and will move the armature to its fully de-energized position.

11

The term tread is still used to refer to the smooth rolling surface of a railroad wheel, even though it has none of the raised features of the tread of a rubber tire.

4-35

Overview of Railroad Signal Circuits

Figure 4-26 Relay Armature and Coil, Safetran Systems Type ST Plug-In Rely (w/o cover)

This is required by FRA rules 234.229 and 236.56. Of course, the track circuit could be adjusted to respond to shunts of higher resistance, and often are, but there are practical limits to this. Because of the less-than-infinite resistance of the track ballast (see Ballast, above), the track circuit must be adjusted so that the conductance of the ballast is not enough of a shunt to deenergize the track relay. Moreover, as the resistance of the ballast changes, in response to the ever-changing amount of moisture in it, the track circuit must be adjusted so that a normal but significant drop in ballast resistance will not falsely de-energize the track relay. Correct adjustment of the track circuit, the use of good-quality ballast materials, and periodic rail and ballast maintenance are all key ingredients in track circuit reliability.

Figure 4-27 Track Circuit Showing Broken Rail

4-36

Overview of Railroad Signal Circuits

Another very desirable characteristic of Robinsons track circuit was its fail-safe, closed-circuit design. Any condition such as a broken rail, loose connection, dead battery, etc., will tend to deenergize the track relay, mimicking occupancy by a train. This feature represents the genesis of the closed-circuit principle in railway signaling, which is required by FRA rule 236.5. Robinsons track circuit quickly underwent necessary refinements such as the placement of bond wires around each joint bar to ensure consistent conductivity where loose joint bars were a problem. Years would pass, however, before truly vital track relays and associated signal apparatus were developed to form a complete block signal system. In 1872, William Robinson had little reason to consider the effects of induced currents or other foreign currents on the performance of his track circuit. Thomas Edisons urban dc utility systems were still 10-12 years away, and the Tesla/Westinghouse ac system did not come into use until 1893. Another common source of track circuit interference, which was not considered in the nineteenth century, was the electric propulsion of trains. In Mr. Robinsons time, this was still science fiction. And other sources of track circuit interference such as impressed potential cathodic protection systems were not even invented until the mid-twentieth century. Perhaps the first hint of things to come was when the insulated joints separating two adjacent track circuits became defective, allowing energy from one circuit to falsely-energize the adjoining track relay. Railroads at this time had no easy means of solving this problem other than to transpose the polarity of adjacent track circuits, and use polarity-sensitive relays. However, this solution cannot be applied where neutral (polarity-insensitive) track relays are used. Track circuits have formed the foundation of almost all railroad signaling today in North America. And the principles of the track circuit have changed little since their inception. But the technology of track circuits continues to evolve. The use of battery-powered direct-current track circuits eventually led to the development of track circuits using alternating currents. Transistors were used in track circuits as early as 1956, and microprocessors were applied to track circuits in 1981. New opportunities for improved track circuit performance and economics continue to appear. Borrowing from trends in modern digital communications, sophisticated encoding and decoding techniques such as trellis modulation, frequency hopping, and digital signal processing are just now being applied to some new signaling equipment. Track circuits of some form will remain in use for many years to come. Current trends suggest that both the number and physical length of track circuits will continue to increase in the future. Railroad History The history of railroad signaling in North America is necessarily intertwined with the history of the railroads themselves. Although the U.S. rail network was being greatly overbuilt by speculators prior to 1900, and peaked at 254,000 route miles in 1916, it presently consists of about 170,000 route miles. However, this decline in route mileage does not accurately reflect the state of the industry, as railroads in the U.S. have seen a four-fold increase in gross ton-miles 4-37

Overview of Railroad Signal Circuits

since World War I. (Ton-miles are computed by multiplying the tonnage of freight transported by the distance moved in miles.) Todays much larger freight traffic load is handled by a small fraction of the trackage, employees, motive power (locomotives), and rolling stock (railroad cars) required during the period from 1880 to 1950, when railroads held a preeminent position in the American economy. One of the keys to this increase in productivity has been the development of signaling systems. General Types of Signal Equipment An overview of many different types of railroad signal equipment will be provided in this section. For the purpose of discussion, we have grouped the equipment into the following three general categories: 1. Signal Systems 2. Grade Crossing Warning Systems 3. Auxiliary Systems Signal systems are systems used to control the movement of trains in a safe and efficient manner. The principal objectives are to keep the trains routed onto the correct tracks, and keep them from colliding with one another or derailing. Signal systems must do all of these things, while also being vital, or in other words: fail-safe, single-fault tolerant, and self-indicating of any vital failure (no latent vital failures allowed). Grade Crossing Warning Systems are the equipment used to detect an approaching train at highway-rail intersections, better known as grade crossings, and provide a warning to motorists and pedestrians. These systems must also be vital. Grade Crossing Warning Systems were once much more thoroughly integrated with the signal systems used in North America than they presently are. Unlike the practices in some other countries, railroads in North America usually employ Grade Crossing Warning Systems that are almost entirely separate from the systems used to control the movement of trains (Signal Systems). No value judgments will be made here, but this difference in engineering philosophy has both beneficial and adverse impacts on the performance characteristics of Grade Crossing Warning Systems worldwide, including their susceptibility to interference from ac power systems. Although some Signal Systems and Grade Crossing Warning Systems in North America are interconnected to some degree, these two types of equipment may be effectively discussed separately with no loss of clarity or understanding. However, as railroad signal systems continue to evolve, it appears likely that we will see an increase in the degree of interconnection between Signal Systems and Grade Crossing Warning Systems in the future. The third category of equipment discussed in this chapter is that of auxiliary systems. For the purposes of this chapter, auxiliary systems are those systems, which may not be directly related to the functions of either Signal Systems or Grade Crossing Warning Systems, but are commonly 4-38

Overview of Railroad Signal Circuits

found on a railroad. Many of them are considered non-vital. In other words, they are not strictly required to be fail-safe, and may not be subject to all of the same federal regulations that apply to signal systems and grade crossing warning systems. However, some of these systems are constructed using many of the same engineering and manufacturing practices as vital signaling equipment. The category of auxiliary systems that we have created here includes both hazard detectors, and commercial tracking systems. Hazard detectors are systems designed to detect the occurrence of specific problems that may affect the safety or efficiency of train operations, such as: landslides, overheated wheel bearings, or dragging cars or other equipment. These hazard detectors may, or may not be, connected directly to the Signal System or Grade Crossing Warning Systems. Some of them are independent, and simply alert railroad personnel or take other action when they have detected a problem. Commercial tracking systems, such as tag readers and track scales, are used to automate the process of tracking goods shipments. Clearly, these systems fall into the non-vital category. But they are nonetheless essential to efficient railroad operations. Some parts of these systems are installed along the wayside of the railroad, some are installed in or near the track, and some are mounted on the locomotives and railroad cars. All of these three different types of equipment can potentially experience interference from ac power distribution and transmission systems, and an attempt will be made here to identify the common susceptibilities of each type. However, particular emphasis will be placed on grade crossing warning systems, as they are generally the first system to experience operational difficulties arising from ac interference. What is a Signal System? Signal systems are railroad signal equipment designed to control the movement of trains in a safe and efficient manner. Modern signal systems usually contain all of the following four items: 1. a fail-safe means of detecting the presence of a train or cars on each portion of the tracks to be controlled by the system (usually done via track circuits), 2. a fail-safe control logic that cannot be easily overridden by a train dispatcher or train crew (regardless of what youve seen in the movies), 3. a fail-safe means of indicating to the trains which movements may safely be made (e.g. wayside signals, cab signals), and 4. a fail-safe means of controlling and detecting the position of each mechanical switch connecting tracks together. Obviously, fail-safe characteristics are an important part of a signal system, and this topic will be given a more thorough examination below. However, at its most basic, a signal system need only contain the first three items. Indeed, the first signal systems constructed had nothing more than the first three, as all of the track switches were operated by hand, and were not electricallyconnected to the signal system at all. 4-39

Overview of Railroad Signal Circuits

Signaled vs. Non-Signaled Trackage Although this chapter is devoted to railroad signaling equipment, less than half of the trackage in North America is equipped with a signal system. Frequently referred to as Dark Territory, traffic in this non-signaled trackage is controlled by various operating procedures, such as: 1. Track Warrant Control (TWC) 2. Timetable & Train Order 3. Manual Block 4. One-Train Operation 5. Line-of-Sight or Flag Protection 6. Train Registers 7. Token Block Some other methods, such as Form D, also exist, and represent a combination of the above techniques. The most common procedure for controlling train movements in dark territory is Track Warrant Control (TWC), which relies on a documented dialogue of train movement directives and train location information between dispatchers and train crews. With the advent of reliable radio communications, TWC is now widely used - and is generally satisfactory for light-to-moderate traffic conditions. Although the Federal Railroad Administration (the FRA a branch of the United States Department of Transportation) regulations do not specifically prohibit it, simple economics dictate that non-signaled territory employing TWC is generally suitable only for lines handling fewer than 15-25 trains per day. The time required for the manipulation of hand-operated track switches by train crews is the primary factor that limits the traffic capacity in non-signaled TWC territory. This handicap has been compounded by the elimination of cabooses on the rear of trains in the early 1980s, which made it impractical for train crews riding in the locomotive to open or close hand-operated track switches behind long trains. For this reason, power-operated, self-restoring switches have been commonplace on some TWC railroad lines since the early 1990s. These switches automatically return to their previous position at some time after the train has left the area, after having been initially moved by the train crew. This arrangement serves as a low-cost substitute for a more sophisticated signal system. However, non-signaled trackage isnt necessarily devoid of all types of signal equipment. There are roughly 265,000 public Highway-Rail Crossings in the United States. Of these, about 65,000 are equipped with some type of automatic Grade Crossing Warning System. Except for a handful of crossings controlled by manual or experimental means, each crossing equipped with warning devices employs some form of track circuit for train detection, just as a signal system does.

4-40

Overview of Railroad Signal Circuits

Many of these crossings are installed in so-called dark territory, which is a term used to refer to railroad tracks which do not have a signal system installed. Under heavy train traffic, a signaling system is the only efficient and cost-effective means of controlling train movements. And the FRA rules specifically require the use of a signal system when freight train speeds exceed 50 miles per hour (M.P.H.), or passenger train speeds exceed 60 M.P.H. (rule 236.0c). Economics have favored the use of faster train speeds, for both passengers and freight, and thus signal systems have become increasingly widespread on U.S. railroads. Signaled Trackage and its Signal Systems There are many different types of signaling systems currently in use to control the movement of trains. They can differ from one another in two primary ways: 1. How (or where) the indications are provided to the train crew (Wayside Signals vs. Cab Signals) 2. The type of control exercised over train movements (ABS, APBS, CTC, CBTC, etc.) Large railroads will usually have some combination of the above, while commuter transit operations tend to be more uniform in their overall signaling scheme. Regardless of signal system type, the safety of most signal systems relies on the train crews compliance with the indications provided by the signal system. However, some signal systems also have the ability to assume limited control of the train in an emergency, usually by activating the trains brakes, if needed. Signal Systems: Wayside Signals and Cab Signals We will begin our discussion of signal systems by starting with the most visible parts. The first parts of a signal system usually seen by a casual observer are the Wayside Signals. As their name implies, Wayside Signals are permanently installed signal equipment situated alongside or above the track, i.e. along the wayside. These signals convey indications by means of visible aspects composed of colored lights, multi-position arrays of lights, flashing lights, moveable semaphore blades, or a combination of these features. The particular style of signal and the set of aspects and indications it can display are determined by the preferences of the owning railroad. Some signals can display only one aspect, while others can display as many as 31 different aspects.

4-41

Overview of Railroad Signal Circuits

Figure 4-28 Rear View of a Wayside Signal Head (Searchlight Type)

A wayside signals aspect is simply its appearance, i.e. what the signal looks like (red, yellow, green, vertical, horizontal, etc.). Each signal aspect conveys specific information (called the indication) to the train crew or operator, and can tell them whether to stop or go, and at what speed, and onto which track the train will travel at the next switch, or even at the next switch beyond that. The indication provided must be consistent and valid during the time that the train is approaching and passing the wayside signal, and is used to control the movement of the train until the train approaches the next wayside signal. Cab signals seek to convey this same type of information to the train operator by transmitting a coded electrical signal through the rails, which is then picked up by sensing coils on the front of the train. The signal aspect is then shown on a small display in the cab, which shall be plainly visible to member or members [sic] of the locomotive crew from their stations in the cab (FRA Rule 236.515). Cab signals are often used in addition to wayside signals, in order to fulfill some of the additional federal requirements for trains operating at 80 M.P.H. or faster in the United States. The primary difference between wayside signals and cab signals lies in exactly how, or more specifically, exactly where, a signal aspect is displayed to the train crew or train operator. Wayside signals are located along the wayside of the track, and cab signals are located in the cab of the lead locomotive. Wayside signals are a form of intermittent control. Once the head end of the train has passed (accepted) the wayside signal, the train crews memory of its indication is all that continues to govern the movement of the train, now that the wayside signal is no longer within view. In contrast, cab signals are a form of continuous control. The signal aspect is continuously presented to the trains operator, and is generally unaffected by poor weather, other visual obstructions, or vandalism. 4-42

Overview of Railroad Signal Circuits

What is a Block? Each signal indication is used to control train movements over a specific physical length of track, known as a block. The ends of a block are often easy to spot, as there will likely be a wayside signal (or several) located at the point where one block ends, and the next one begins. However, the FRA has a more specific definition: 236.708 Block (From FRA Rule 236, Subpart G: Definitions): A length of track of defined limits, the use of which by trains is governed by block signals, cab signals, or both. Although we have been discussing wayside signals, and the FRA definition refers to block signals, the terms may be used interchangeably in this case. By definition, block signals are signals located at the entrance to a block, which display their aspects in one direction only, so that they can be seen by trains before they enter that block. Signals will often be constructed in a back-to-back fashion, often on the same signal mast or signal bridge, so that signaling can be provided for trains running in either direction along the track(s). Another feature of the track that will usually accompany the end of a block (and a wayside signal) is the presence of insulated joints. These are special rail joints that physically connect two pieces of rail together, while still keeping them electrically insulated from each other (see Insulated Joints, above). The FRA definition of a block does not explicitly mention insulated joints, because not all signaling systems require them (Audio-Frequency Overlay track circuits, for example).

4-43

Overview of Railroad Signal Circuits

Figure 4-29 Section vs. Block

The term block is sometimes loosely applied to individual track circuits separated by insulated joints, but this is not correct in the strictest sense. In signal systems, a signal block is a logical entity, and a single signal block may contain several discrete track circuits, separated by insulated joints, whose outputs are taken as a whole to determine the occupancy of the block (i.e. are any of the track circuits in the block occupied). Signaling Basics In the most basic form of block signaling, Green, Yellow, and Red signal aspects convey indications of Proceed, Approach, and Stop, respectively. Other aspects, or combinations of aspects, can provide more detailed speed limit information, or indicate the position of switches in the track ahead (mainline vs. siding or diverging route). Like highway traffic signals, Green indicates permission to proceed at a rate not exceeding the applicable speed limit, and Red means stop, but Yellow is slightly different. For automobiles, yellow means that the specific traffic light that you are looking at is about to turn red. For trains, yellow means that the next wayside signal ahead presently isnt green, and may either be red or yellow. Yellow signals instruct trains to slow down (usually below 35 M.P.H.) in preparation for a possible stop at the next wayside signal. 4-44

Overview of Railroad Signal Circuits

And just like a dark traffic light, a dark wayside signal is an indication to stop. This is one of the rules required to make the system vital (more on this topic below).

Figure 4-30 Limited Clearance for Signals and other Equipment on a Rapid Transit Line.

The location of wayside signals is based on many factors, including operational requirements, braking distances, sighting distance, availability of power or communications, and other factors such as available space (see picture, above). As a general rule, wayside signals are located wherever it may normally be necessary and desirable to stop a train. What is an Interlocking? Dictionary definitions for the term interlock often include the railroad sense. But this definition is far from complete, as many important details are lacking. Here we will attempt to explain the basic principles of railroad interlocking, and some of the associated terminology. The terms interlock and interlocking are used interchangeably as nouns in the railroad sense, with the term interlocker being the slang or colloquial form. All of these terms refer to the machinery of an interlocking, and not to the people using it. The FRA has official definitions for the term interlocking which refer to things which must succeed each other in proper sequence (rules 236.750, 236.751), and associates this with the behavior of the signal system and its associated track switches. In brief, it all boils down to this: the signal aspects must be displayed, and the track switches must be moved, in such a fashion as to prevent trains from running into each other, or running off the tracks. Much of the care and craft put into signaling systems by signal designers goes toward satisfying these essential requirements.

4-45

Overview of Railroad Signal Circuits

Most of the specific hazards that interlocks avoid revolve around track switches. Switches, whether manually or remotely operated, are the mechanical devices that allow two separate tracks to merge into a single track, or allow one track to diverge into two. In either case, only two of the three (or more) tracks that come together at a switch can be connected together at any one time. Thus, switches serve as the intersections of merging or diverging railroad tracks. The positions of track switches must be carefully controlled, and must be coordinated with the signal indications given. Two trains must never both be given green lights to merge onto the same track simultaneously, even if they are heading in the same direction. If a train is currently occupying a switch, that switch must not be allowed to change positions. And once a train has been given permission to pass through a switch as the train proceeds from one track to another, any trains on the third track connected to the switch must be stopped before reaching the switch. This requires that they be given enough advance warning to allow them to safely brake to a stop at a safe distance from both the switch and the train currently using it. This means that the interlocking (and signal system) must have the ability to control the display of signal aspects that may be shown on wayside or cab signals that are miles away from the switch itself. When switches are placed near one another, to create the sort of switching network frequently needed in switching yards, rail terminals, or at crossovers in multiple-track territory, the complexity of the signaling task appears to rise exponentially. Indeed, this is true. At any given time, the allowable changes at each signal or switch may be dependent on the state or condition of any or all of the other signals and switches in the area, as well as on the location(s) of any train(s). Where such an assembly of track switches exists, as is common in switching yards and rail terminals, an interlocking is provided. Unlike some other types of signal systems discussed below (e.g. ABS and APB), train movements into and through an interlocking are controlled solely by signal indication. Interlockings may be locally, remotely, and/or automatically controlled. A mechanical or electrical interlocking of the switches and signals ensures that conflicting routes and other unsafe conditions cannot result from improper manipulation of the controls by dispatchers (remotely), or by train crews (locally). In addition to the control of signals, switches, and derails, interlocked functions may include moveable bridges, tunnel doors, moveable catenary couplers and other special devices.

Figure 4-31 Elementary Interlocking Layouts

4-46

Overview of Railroad Signal Circuits

Mechanical Interlocking Mechanical Interlocking (developed 1870-1900) used mechanically-interlocked, manuallyoperated levers in the signal tower that were connected directly to switches and signals by steel cables, or 3/4" or 1" diameter steel pipes and rods supported on greased rollers. The interlocking mechanism thus created embodied the necessary control logic to prevent the states of signals and switches from being changed in an improper sequence. Due to the mass and friction of the mechanical interconnections, the mechanically operated switches and signals controlled by the interlocking operator could be located no more than about 1500-2000 feet from the signal tower. Electro-mechanical locks and slotting mechanisms were later added to the mechanical levers, to improve safety and flexibility by making the system interactive with train movements. That is, the mechanism could now sense the presence of a train on a particular section of track via an input from a track circuit. The last new mechanical interlocking was installed in the United States during World War II. As of 2003, the few surviving mechanical interlocking plants in the U.S. will be retired in the near future. Worldwide, a substantial number of mechanical plants still exist, primarily on lines of lesser importance, or where economic conditions have inhibited modernization. The production of spare parts by the last British mechanical interlocking manufacturer ceased in the early 1970s.

Figure 4-32 Mechanical Interlocking Machine

Power Interlocking Power Interlocking technology employing hydraulic, electro-pneumatic, and all-electric field equipment was introduced beginning 1890-1900. As in earlier mechanical interlocking systems, control levers for switches and signals were mechanically and electrically interlocked to prevent improper manipulation. The maximum control distance was substantially increased, to approximately 5,000 to 10,000 feet from the control machine. This technology matured prior to World War I, with new installations as late as 1955. One of the largest machines of this type was delivered in 1949. Although rapidly disappearing from U.S. railroads, a substantial number of power interlocking machines may continue in-service for another 10-20 years on certain rapid transit lines. 4-47

Overview of Railroad Signal Circuits

Figure 4-33 Typical Power Interlocking Machine

Relay-Based Interlocking Relay Interlocking, introduced in 1929, used fail-safe Vital Relays to substitute for the cumbersome and inflexible mechanical locking bars and levers of mechanical interlocking systems. The positions of the track switches, as well as track occupancy, were all represented by relay states. This is very similar to Power Interlocking, above, but with no mechanical interlocking. With over 4000 of these vital relays, the largest relay interlocking plant installed in the U.S. was placed in service in 1994. Interlockings, grade crossing warning systems, and signal systems employing these same vital relays continue to be widely produced, and will remain in service for many years to come.

4-48

Overview of Railroad Signal Circuits

Figure 4-34 Rack-Mounted Plug-In Vital Relays in a Relay-Based Interlocking Plant

Solid-State Interlocking Solid-State Interlocking (S.S.I.) technology was introduced in the U.S. in 1985 as a substitute for the large number of costly vital relays found in most interlockings. Solid-state interlockings employ microprocessors in combination with special I/O circuitry and software functions that ensure prompt detection of, and proper reaction to, any internal failure.

Figure 4-35 Sophisticated Electronics Typical of Solid-State Interlocking Systems

4-49

Overview of Railroad Signal Circuits

Types of Signaling Systems Over the history of railroad signaling, there have been many different prevailing philosophies as to what constituted the optimal signal system for a particular application. This process of signal evolution is continuing even today. But due to the installed cost and long lifetime of a signal system, the process of updating signal systems has continued in a piecemeal fashion, as dictated by the economics of each situation. As a result, there is still a wide variety of signal systems installed on the signaled trackage of the North American rail network. This section briefly discusses the most common. In order to organize these systems by their features, we will begin with the most visible difference between them, i.e. the method of displaying the signal aspect, and then move into the more conceptual differences between railroad signal systems. There are currently two basic and distinct ways of displaying a signal aspect or indication to the crew of a train. The first, and oldest, is by wayside signals installed alongside the tracks. The other is via a display in the cab of a locomotive or train. We will begin with wayside signals. Wayside Signal Aspects and Indications Wayside signals can take many forms. Many obsolete systems using colored balls, flags, and other innovative signal equipment have been used in the past. However, current FRA rules permit only the use of semaphore blades or lights to create signal aspects and indications (FRA Rule 236.23). And even though they are still legal according to FRA rules, the use of semaphore arms has been almost completely discontinued in the United States. The colors used for signal aspects today are generally limited to green, yellow, red, and lunar white, which is a bluish-white. A truly blue signal light has a special meaning. The presence of a blue signal invokes an extensive list of rules and regulations designed to protect the safety of people working on and around tracks, rolling stock, and on-track machinery. Therefore, the use of blue signal lights is reserved for marker lights only, as are used to comply with FRA rule 218. This rule requires the use of blue signal lights to identify track, cars, track equipment, or locomotives currently being serviced by railroad employees or other persons. Noviol, a fog-penetrating yellow color, is used for position-light signals. Position-light signals, as opposed to colorlight signals, convey their aspects not by the color displayed, but by the orientation or position of a line of lights of the same color. This line emulates the visible position of the formerly used semaphore arms or blades.

4-50

Overview of Railroad Signal Circuits

Figure 4-36 Common Types of Light Signals

Figure 4-37 Track Plan Symbology for Wayside Signals

A home signal is a signal located at the entrance to a route or block (i.e. at a switch), which governs the movement of trains into that route or block. Obviously, by this definition, almost every signal could be called a home signal. However, this term is usually reserved for signals located at a switch, with signals located along tracks where there are no switches generally being called intermediate signals. 4-51

Overview of Railroad Signal Circuits

Two-arm and three-arm (two-head and three-head) interlocking home signals are common. The aspect displayed on a signal is normally read from top to bottom. A signal displaying Greenover-Red-over-Red (G/R/R) indicates clear - proceed at maximum authorized speed. The G/R/R aspect gives the same indication as a two-arm signal displaying G/R, or a single-arm G. However, the multiple-arm signals can also provide other information to the trains crew about the route that they will be taking over the next few switches in the track ahead.

Figure 4-38 3-Aspect and 4-Aspect Signal Systems

Colorlight Signals Presently, commercial trends suggest that the 3-lamp colorlight signal is the most versatile and economical form of wayside signal in use. Their light units may be arranged in a vertical array, with an oval background, or in a triangular form, with a circular background, as shown below.

Figure 4-39 Colorlight signals (Safetran Systems)

4-52

Overview of Railroad Signal Circuits

Figure 4-40 Colorlight Signals

Although controlled exclusively by vital relay logic in the past, the 10-volt, 18-watt incandescent lamps most often used in colorlight signals are readily controlled using modern vital solid-state output devices. Unlike searchlight signals (see below), no indication circuits or stuckmechanism checking is required for this type of signal. One optical device conspicuously absent from the design of colorlight signal heads is a mirror. Colorlight signals having focused single or doublet lenses in them do not have mirrors behind their lamps. Only a small fraction (less than 20%) of the light output of the filament is captured and projected through the lens system. The rest is absorbed by the interior of the signal assembly, which is usually painted flat black. Mirrors cannot be used to capture a greater percentage of the incandescent lamps output in most colorlight signal head designs, due to the possibility of phantom indications. Phantom indications are caused by the reflection of light from external sources such as sunlight or locomotive headlights. Light entering an optical system with a mirror could potentially reflect outward again, passing through the colored filter or roundel, and giving the impression that the signal is lit in a particular color, even when it isnt lit at all. Another technique for eliminating phantom indications is the installation of a louvered screen or grating in front of the lens. These are not unlike the screens sometimes used on highway traffic light to restrict the angles at which the signal can be viewed. In some cases of reported phantom indications, it has been impossible to duplicate the reported failure, and it is unknown whether the false indication was caused by a phantom, or by induced ac voltage on the lighting circuit wires. But phantom indications are taken very seriously by signal design and maintenance departments. Crossing Flashers In contrast, the large (8-3/8" or 12" diameter) flashing incandescent lights used at grade crossings do have a concave a mirror behind the 10-volt, 18- or 25-watt lamp. This greatly 4-53

Overview of Railroad Signal Circuits

increases their light output, and their visibility. Phantom indications are not a problem in this case, as a phantom indication to a motorist or pedestrian could only result in their stopping briefly for a train that wasnt there. Therefore, these lights can also be equipped with a colored roundel that distributes light over a relatively wide angle, and no louvered screens are use. However, they may be equipped with a wire mesh to reduce damage from vandalism.

Figure 4-41 Flashing Lights from a Grade Crossing (Shown Without Visor or Background) (Safetran Systems)

LED Colorlight Signals The use of Light-Emitting Diode (LED) light units in colorlight signals is currently increasing. These light units have been made by installing a large number of LEDs into a printed circuit board, forming a two-dimensional planar array. These resemble the LED light units now commonly installed in highway traffic lights.

Figure 4-42 LED Wayside Signal Lamps (Safetran Systems)

4-54

Overview of Railroad Signal Circuits

LED lamp unit manufacturers have now successfully mimicked the standard red, yellow, and admiralty green colors used in signaling. The use of colored filters in conjunction with colored LEDs is unnecessary, and even inadvisable in some cases. White LEDs have not yet been successfully demonstrated as a direct substitute for incandescent lamps, but this could still happen. As each LED is molded into its own plastic lens, these light units have the distinct manufacturing advantage of not requiring precise focusing, as was required for the doublet-lens optical system used with incandescent lamps. Searchlight Signals Another longstanding type of wayside signal is the searchlight signal. These consist of a single bulb placed behind a lens assembly and color-selection mechanism. This mechanism controls the aspect (color) displayed by electro-mechanically moving the correct colored lens into position in front of the bulb. Introduced in 1920, the searchlight mechanism projects a beam of light through one of the colored filters mounted on a moving spectacle. The position of the spectacle is controlled by an electromagnetic mechanism driven by low-current dc energy. Because stray light entering the searchlight mechanism from some external source cannot easily produce a phantom indication, virtually all of the lamps light output may be captured by an ellipsoidal reflector. And since the mechanism is polarity-sensitive, this usually limits their ac susceptibility. Today, the chief advantage of the searchlight signal lies in its ability to produce a clear, distinctive indication with the sun at low angles behind or facing the signal. The searchlight signals carefully aligned optics provide an intense beam of colored light that can be precisely directed only at trains on the track associated with that signal. This is very helpful in multipletrack territory where there are several tracks running in parallel to one another. A clear association between a signal and the track it governs is generally required by FRA Rule 236.21.

Figure 4-43 Rear View of a Wayside Signal Head (Searchlight Type)

4-55

Overview of Railroad Signal Circuits

Fiber-Optic Searchlight Signals Other wayside signal designs have used separate incandescent bulbs, each with its own colored filter to create colored light that is carried by fiber optics to a single lens assembly. The bundled ends of the fibers are positioned at the focal point of the lens assembly, just as the filament of an incandescent bulb would be. These signal heads are controlled in exactly the same way that ordinary colorlight signals are. The primary advantage of this design is that it combines the colorlight signals simplicity of control with the compactness of a searchlight signal. In some cases, this allows it to be installed in unusually tight spaces, and is currently the only way to display four different colors through a single lens.

Figure 4-44 Safetran Unilens Signal (Safetran Systems)

Position Light Signals Although physically larger and more complex than simple colorlight signals, large numbers of position light signals on former Pennsylvania Railroad and Norfolk &Western trackage in the eastern U.S. will remain in service for many years to come. Many of these position light signals have now been modified to some form of hybrid color/position signaling. The distinctive and versatile Baltimore & Ohio-style color-position light signals found on former Baltimore &Ohio, Illinois Central, and Gulf Mobile & Ohio railroad lines are gradually being replaced by two-arm colorlight signals. Position light signals continue to find use in other parts of the world, however. Semaphore Signals As of 2003, most of the remaining motor-operated semaphore signals in-service on major U.S. railroads were quickly being eradicated, partly due to lack of spare parts to replace those removed (stolen) by collectors. Although exceptionally long-lived, the semaphore signal has no inherent technical advantage in any application, and was widely considered obsolete by 1940.

4-56

Overview of Railroad Signal Circuits

Signal Lighting Circuits Low-voltage (8-10 volt) lighting circuits have remained the standard for both signal and grade crossing light units for decades. Power for almost all lighting circuits in railroad signaling is carried by heavy-gauge cable (#6 A.W.G. or larger) from the point of control to the signal head. Due to I2R losses at such low voltages, most cable runs are limited to less than 250 in length. The cable runs of transformer-coupled ac lighting circuits, which have a step-down transformer in the signal head, can go much further, but are still typically limited to about 3500 in length. Although the use of such low voltages might seem problematic, these low-voltage lighting circuits offer several distinct advantages over higher-voltage alternatives. First: rugged, compact lamp filaments are harder to fabricate for operating voltages greater than 12-24 volts. Second, in a low-voltage system, the difference in current between the on and off states of the lamp is very great, enabling the use of very simple filament failure detection circuits. Thirdly, the lowimpedance, high-current nature of these circuits makes it more difficult to falsely energize them with stray energy. Lighting circuits employing the simplistic 1st-generation LED-based replacements for incandescent bulbs, or other circuits operating at lower currents and higher voltages than incandescent bulbs are inherently more susceptible to EMC problems, especially where there is significant capacitive coupling to other circuits in long multi-conductor cables. The relatively high current required to energize low voltage incandescent lamps also reduces the risk of them being falsely energized due to unintentional grounds, crossed wires, or magnetically induced energy. Signal Lamps Most wayside signals use incandescent lamps with 13 to 25 watt, 10-volt filaments, with the 18-watt and 25-watt sizes being the most common for outdoor applications. These lamps resemble the common #1141 automotive tail lamp bulb available at your local auto parts store, but are of a special design requiring meticulous construction (and a premium price). This is due to the need for a small, concentrated point source of light. The position of the coiled coil tungsten filament of the lamp is carefully adjusted during manufacture to position it within 0.016 inches of the optical systems focal point. Rated for 1500 hours at 10 volts, these lamps have a nearly unlimited life when operated at 8.0 to 8.5 Volts (ac or dc), as is the usual practice. Lamps as small as 5 watts are used in subway tunnels, where contrast with high ambient lighting conditions is not a problem. These may also be used for approach-lit searchlight signals (above ground) that are fed from primary batteries, in order to conserve battery power. Dual filament lamps are common in many countries outside North America. Filament detection relays are located in the signal head, and often provide a filament failure indication to the interlocking or Central Traffic Control (CTC) operator when the first filament burns out. Operation is transferred to the secondary filament when this occurs. Dual filament lamps having 13-watt or 18-watt major filaments, and 3.5-watt minor filaments, are standard for some U.S. railroads as well. However, their use is generally declining. 4-57

Overview of Railroad Signal Circuits

Lamps having two identical half-wattage filaments are sometimes used for railroad signaling to provide a graceful degradation of performance when one filament fails. However, as with any dual filament lamp, it is impossible to locate both filaments at the exact focal point of the optical system. The out-of-focus secondary filament tends to produce a much less visible signal indication. Fortunately, FRA rules for failed lamps require only that the signal not cause the display a less-restrictive aspect than intended as the result of a failed lamp12. The FRA has also interpreted this rule to mean that a multi-arm signal (a wayside signal with two or more signal heads) may not be deliberately made completely dark as the result of a light-out condition. Cab Signal Systems All of the signal heads discussed to this point are used in some form of wayside signaling, where the signal indication is conveyed to the train crew at fixed locations along the tracks. In contrast, cab signals show their colored aspects or numeric speed limit indications on a display inside the cab of the locomotive. Cab signals remain clearly visible to the train crew even when wayside signals are obscured by track curvature, fog, or other hindrances. The various forms of cab signaling used in the U.S.A. today were introduced between 1923 and 1928. In most instances, cab signaling serves as an adjunct to the wayside signaling system. In one of its more sophisticated forms, cab signaling provides the basis for Automatic Train Control (ATC) (see below). As with wayside signals, the cab signal system provides information to the train, based on block occupancy and route conditions immediately ahead. Unlike a wayside signal system, the cab signal system furnishes a continuous stream of real-time information as the train traverses the block. Any change in indication, whether more permissive, or more restrictive, is immediately presented to the train operator. Cab signaling systems work by transmitting a current through the rails, which is sensed by coils mounted on the front of the train. Types of Signal Control Systems Regardless of whether a signal systems uses wayside or cab signals to display a signal aspect to a trains crew, and thereby direct the movement of trains, it must still have the required vital logic to control these aspects and indications. There are several different styles of train control, all of which are vital, but they differ slightly in the degree and type of human involvement. Automatic Block Signals (ABS) At its most basic, an ABS system simply tells a train crew about the occupancy and condition of the tracks ahead on a single track. If there were only one train on an all-ABS railroad with only one track and no switches, the signals ahead of the train would always be green, unless there was
12

FRA rule 236.23, subpart f.

4-58

Overview of Railroad Signal Circuits

a problem with the tracks or track circuits themselves. In that case, the ABS system should show a yellow signal two or three blocks before the train arrived at the block with the problem, and would show a red signal at the entrance to the block with the problem. As soon as we add the complexity of switches and multiple routes or tracks, things change. Specifically, the first time the train on our one-train railroad encountered a switch that wasnt lined up for the train to continue on the main line, the signal for the main line in front of that switch would not show a green aspect. This is because the train would be forced by the switch to diverge onto another track, and would therefore not be proceeding at the timetable speed limit along the main line. All of this can be automatically accomplished by simple track circuits, wayside signals, vital relays, and switch circuit controllers. No human intervention is required in order for the signal system to provide an indication to the train, hence the word automatic in the name. The simplest Automatic Block Signaling systems provide a means of safely expediting successive following train movements, in one direction at a time, on mainline track. Of course, ABS systems are invariably arranged to provide bi-directional running capability. However, trains can still only use it in one direction at a time. In a typical ABS system, the track is divided into a series of consecutive blocks of varied length, 13 based on topography and operating requirements . ABS systems on parallel tracks are generally independent of one another, except for common elements such as power sources and equipment housings. Entry into each block is governed by a wayside signal, the aspects of which are automatically controlled solely by the presence (or absence) and direction of preceding trains. Additionally, block signals usually provide information regarding the aspect displayed on the next signal beyond the one currently being viewed. Block signals do not convey information regarding permanent speed limits, which are instead listed in the railroad timetable. Block lengths are seldom completely uniform. Arbitrary one-mile blocks were common in many older ABS systems since this distance was usually about the maximum operable length for early dc track circuits. Today, individual block lengths on mainline railroads generally range from 4,000 feet to over 15,000 feet, depending on operating requirements. ABS systems use conventional track circuits of many kinds, and often use long cables or open wires on poles to convey the occupancy information along the length of the railroad from block to block. The susceptibility of ABS to ac interference depends on the specific type of track circuits used, and the method used for conveying the occupancy information. Open line-wires will generally be more susceptible to induced ac interference than shielded cables.

13

The terrain through which a railroad runs has a significant impact on braking distances, with trains traveling downhill needing longer to stop.

4-59

Overview of Railroad Signal Circuits

Figure 4-45 Sequence of Aspects for Typical Single Direction Automatic Block System (ABS) in Multi-Track Territory

Absolute-Permissive Block Signaling (APBS) A later improvement (circa 1911) in ABS technology was Absolute-Permissive Block (APB). APB is a somewhat more sophisticated application of ABS to single-track lines where both opposing and following train movements must be controlled. Using relay logic, the APB system is capable of controlling traffic based on the direction of preceding trains. Outside of CTC territory, ABS and APBS signals serve only as adjuncts to movement authority control procedures such as Track Warrant Control (TWC). A distinguishing feature of ABS and APB is that train movements are controlled solely by the location and direction of train movements. There is generally no means of dispatcher control or intervention in controlling signals or switches. All signals normally display their most permissive aspect.

Figure 4-46 Bi-directional Absolute-Permissive Block System (APB)

4-60

Overview of Railroad Signal Circuits

Figure 4-47 Overlap Versus Non-Overlapping Block Systems

Centralized Traffic Control (CTC) The desire to use remotely controlled track switches in outlying areas led to the development of small-scale remote control systems built using telephone selector switches by the early 1920s. And the substitution of vital relays for mechanical interlocking logic was another early step toward the remote control of track switches. The introduction of the compact plug-in vital relays by 1934 also greatly improved the technical and economic feasibility of CTC, because of the smaller physical size of the relays, which were used by the dozens. At that time, a clear distinction developed between devices, circuits and systems that directly affect the safety of train movements (Vital Circuits), and elements whose malfunction could not jeopardize safety (Non-Vital Circuits). Previously, all of the circuits associated with interlocking and ABS logic were vital circuits. With the development of CTC, the concept of non-vital signal circuits was introduced. The distinction between vital and non-vital function is one of the most important concepts in modern railway signaling design. An important result of the clear separation between vital and non-vital circuits in CTC is that the control and indication information that is passed back and forth between the CTC dispatching office and the signal equipment in the field is non-vital. Furthermore, the control logic is implemented entirely using vital devices such as vital relays in located in wayside equipment shelters. Thus the vital state machine constructed may be distributed over hundreds of miles of track, and has one very important feature. Because the vital logic is in the field, and will not permit an unsafe sequence of manipulations by a dispatcher, non-vital communication circuits can be used to control it. Should an unsafe request be sent out by the CTC office, whether due to human error, criminal intent, or communications link malfunction, the vital logic in the field will not allow it to be acted upon in an unsafe manner.

4-61

Overview of Railroad Signal Circuits

Figure 4-48 #20 Switches Allow Crossover Moves Between Tracks at 45 mph in CTC Territory. This Greatly Enhances the Speed and Efficiency of Train Movements

From 1980 to 2000, the number of miles under CTC control substantially increased, due in part to the advent of large computerized dispatching centers. Overall, the number of signaled track miles has increased, while the overall rail network shrank (in route miles) during the same period by 35%.

Figure 4-49 Typical CTC System

Under CTC operation, all train movements within a specific territory are controlled by signal indications. A CTC system is essentially an enormous interlocking, spanning a wide geographic area. Using electrically-operated field equipment, modern communications technology, and nonvital remote control systems, there is no practical limit to the physical size of a CTC system, nor any limit as to the distance between the CTC control center and its field locations. 4-62

Overview of Railroad Signal Circuits

An interlocking, as defined earlier, that is located within CTC territory, is termed a Control Point. Between control points, ABS signals are provided as needed in order to expedite following train movements and provide an indication of the route established (mainline or diverging) over the next turnout or crossover. To encompass the various names applied to different forms of CTC, the FRA uses the term Traffic Control System (TCS) to describe any system whereby: 1. all train movements are controlled by signal indication, and 2. TCS operators have no means of visually observing the head and tail ends of passing trains to confirm that specific trains have entered or vacated a block.

Figure 4-50 Complex Interlocking Control Panel on Heavy Rail Rapid Transit System

Figure 4-51 Computer-Based Operations Control Center for a Light-Rail Transit System (U.S.&S Inc.)

Train Control Systems In addition to developments in ABS, Interlocking, and CTC, another distinct class of signaling technology evolved during the 20th century. These Train Control Systems include Automatic 4-63

Overview of Railroad Signal Circuits

Train Stop (ATS), Automatic Cab Signal (ACS), and Automatic Train Control (ATC) systems. Briefly stated, these systems are overlaid upon ABS and interlocking systems, and serve to enforce compliance with restrictive signal indications. ATC systems may also serve to restrict train speeds in accordance with prevailing civil speed limits. Basically, these systems automatically force the train to slow down or stop, as indicated by the signal system. The most sophisticated train control systems implemented to date offer Automatic Train Operation (ATO) capability. ATO systems provide a means of automatic or semi-automatic train operation, including precise stopping at station platforms or loading/unloading facilities. ATC/ATO systems designed for driverless freight and passenger operation were placed in regular service as early as 1962 in the U.S.A. and Canada. Today, driverless ATC/ATO systems on steel wheel/rail rapid transit lines provide 90-second headways (the spacing between successive trains traveling on the same line) with station stopping position accuracy of better than +/-9". Mechanical Trip Stop The earliest practical train control systems (circa 1900) employed mechanical trip stop devices located at each wayside signal. When a signal displayed its most restrictive aspect, the trip arm was raised or otherwise positioned to make contact with an arm attached to a valve or switch on the passing train. Engagement of the car-borne trip cock placed the trains air brakes into the Emergency mode (the most drastic and rapid form of braking possible, which can result in derailments). Despite the various weaknesses of this system, it remains in-service on a substantial number of rapid transit properties in North America and elsewhere.

Figure 4-52 Mechanical Trip Stop, Wayside Element

4-64

Overview of Railroad Signal Circuits

Figure 4-53 Mechanical Trip Stop, Carborne Element

Intermittent Inductive Trip Stop Mainline railroads operating at high speeds over long distances could not tolerate the handicaps inherent to the mechanical trip stop system. Responding to Interstate Commerce Commission (I.C.C.) orders in 1922 and 1924, requiring the installation of train control systems on certain mainlines, the Intermittent Inductive Automatic Trip Stop system was developed and widely deployed in the U.S.A. by 1930. The wayside element (inductor) located at each signal served to transmit a single bit of information (clear or danger) to passing trains. A wayside inductor (usually mounted in the middle of the track), wound around a C-shaped core, was short-circuited when the associated wayside signal displayed its most permissive aspect. The coil remained open when any other aspect was displayed. The locomotives carried a similar C-shaped core having two coils; one coil providing a steady magnetic flux, while a secondary coil sensed the voltage induced when passing the wayside element. With the wayside coil shorted, the voltage induced in the trains secondary coil was minimal. With the wayside inductor circuit open, the induced voltage was sufficient to trigger the toggling of a relay in the on-board system. This action initiated the application of the trains brakes unless there had been a manual actuation of a forestalling switch by the locomotive engineer. The carborne and wayside equipment were relatively simple, rugged, and could be readily overlaid onto existing signal systems and locomotive brake valves with minimal effort. Nonetheless, the intermittent inductive system had several serious weaknesses, and its use was largely discontinued (with FRA approval) in the early 1970s. Since then, several extensions to the few remaining systems have been made, and one new system of this type has recently been installed on a new freight/passenger light-rail line. 4-65

Overview of Railroad Signal Circuits

Indusi A variation on the intermittent inductive system known as Indusi is widely used throughout the German DB-affiliated railways of Europe. In this system, the wayside element is tuned to differing frequencies (500, 1000, and 2000 Hz). The train detects the presence and frequency of the wayside element, and then governs train movement accordingly. Introduced in the early 1930s, this system has undergone various evolutionary enhancements, including wayside-to-train data transmission capability. Magnetic Trip Stop Various types of intermittent magnetic trip stop systems have been developed in the U.S.A. and Europe since the 1920s. In general, magnetic trip stop systems have a wayside element that produces one or more magnetic fields. The car-borne element senses the polarity and sequence of these fields as the train passes. This type of system may be arranged to provide directional bias, i.e. the wayside element will act upon trains in one direction only. Using a timer to measure the elapsed time between passing successive wayside elements, this system may also be used to control train speed. Magnetic trip stop systems have been applied to a number of newly constructed electrified light rail systems in the U.S. during the past 25 years. Because this type of system is susceptible to stray polarized magnetic fields of a constant polarity, special measures must be taken to avoid false tripping where heavy dc propulsion currents are used.

Figure 4-54 Magnetic Trip Stop System

4-66

Overview of Railroad Signal Circuits

Figure 4-55 Modern Driverless Automatic Train Control Systems Promote Efficient Land Utilization in Densely Populated Urban Area (U.S.&S Inc.)

Communications-Based Train Control (CBTC) CBTC is generally synonymous with a variety of acronyms and trade names: CBS: Communication-Based Signaling TBS: Transmission-Based Signaling PTC: Positive Train Control PTS: Precision Train Control AATC: Advanced Automatic Train Control ARES: Advanced Railway Electronics System ATCS: Advanced Train Control System ETCS: European Train Control System ERTMS: European Rail Traffic Management System

Since the early 1980s, a series of initiatives have been undertaken by various entities in North America and Europe to develop and deploy vehicle-borne train control systems that would augment or eliminate reliance on wayside-oriented signaling and train control technology. At first glance, this would appear to offer a comprehensive solution to power line EMC problems.

4-67

Overview of Railroad Signal Circuits

However in the face of daunting technical and financial hurdles, none of these concepts has yet come to fruition in the U.S. or elsewhere. A long succession of proponents have promised but failed to convincingly demonstrate that any new train control technologies will increase hauling capacity or reduce point-to-point transit times. Communication-based train control systems installed on a limited number of transit lines to-date have tended to be experimental, one-of-a-kind systems having unsatisfactory service histories and short life cycles. In several instances, the product development cycles have been so long that key pieces of hardware, such as the radio transceivers, have become obsolete before initial trials were completed. The technological spectrum of CBTC is divided between loop systems, and wireless systems. First developed in Europe in the late 1960s, loop-based systems such as LZB and SELTRAC employ a continuous or semi-continuous circuit placed parallel to the track with one wire against a running rail and the other wire centered between the running rails. Other types of loops may be in the form of a pair of wires or, in some recent installations, a leaky (i.e. lightly-shielded) coaxial cable. Some loops may extended unbroken for several kilometers or more. Information is passed between the train and the signal system via these loops. The effect of longitudinal ac induction on loop-based systems is a significant EMC concern. More recently, CBTC systems have been developed using wireless communication technologies. Most utilize GSM-R or a proprietary cellular WAN concept. A 2.4 GHz system closely resembling the IEEE 802.11b wireless digital communications standard has been adopted by New York City for CBTC systems. Elsewhere, San Franciscos BART has installed an experimental position-determining and communication system based on military EPLARS spread spectrum radio technology. To date, none of these technologies has successfully supplanted the existing signal system. Track Circuits in General Most Signal Systems and Grade Crossing Warning Systems rely primarily on vital track circuits of one type or another for train detection. If we look at track circuits in general, we will discover that they all share some common features. These features will be present, in one form or another, in all of the different types of track circuits that will be discussed below. They are: 1. a source of electrical energy (e.g. a battery) 2. a means of detecting that electrical energy (e.g. the coil of a relay) 3. a means of transmitting the energy from the source to the detector in such a fashion that it can be changed in some way by the presence or location of a train (i.e. the rails and the axles of the train) 4. an output which provides information about the presence or location of a train to other components of the railroad signal system. (e.g. the armature and contacts of the relay, a relay-drive output, etc.)

4-68

Overview of Railroad Signal Circuits

Batteries and Electrification Even though this handbook is focused on railroads and ac power lines sharing the same corridor, this practice was once much less common than it is today. As such, the availability (and reliability) of early ac power to power the signaling equipment was a serious concern for the railroads. For many years, railroads relied almost exclusively on primary batteries, or even used their own generation and distribution facilities. Every track circuit relies on a source of electrical energy that can be converted into the proper form for transmission into or along the tracks. In modern times, the ultimate source of that electrical energy is usually the local electric utility. Due to their need for extreme reliability, railroad signal systems normally use rechargeable batteries, kept fully charged by an ac-powered battery charger. The chargers used are large enough in capacity to supply the ongoing needs of the signal equipment, and can simultaneously provide a fairly rapid recharging of the battery after a deep discharge. In the case of simple dc track circuits, which apply direct current to the track at roughly 1.6 volts, the source of energy is usually a single rechargeable lead-acid cell kept charged by a battery charger. The battery charger is normally powered by 120 Volt or 240 Volt 60 Hz ac power, but in remote installations may also be supplied by an array of photovoltaic solar cells, wind generator, or even a thermoelectric generator. Other more sophisticated track circuits are often powered by a similar 12 Volt battery, whose state of charge is maintained in much the same way. Modern track circuits are no longer restricted to outputting a simple dc current derived directly from a battery, and many different types of pulsed-dc or ac electrical signals are now applied to the tracks. But since the used of direct-current power at 12 Volts with battery back-up is the most common form of power supply for railroad electronic signaling equipment, most transmitted signals are produced by track circuits that are somehow battery powered. A key catalyst in the switch from primary batteries to rechargeable batteries and ac-powered rectifiers was the development of a suitable rectifier. Earlier rectifier technologies had proven unsuitable, but with the development of the first solid-state (copper oxide) rectifier in 1926, reliable, low-cost battery chargers became available for use. Since that time, this technology has been successively replaced by selenium rectifiers, silicon diodes, and now microprocessorcontrolled temperature-compensating battery chargers. But it was the development of a suitable rectifier that made possible the use of rechargeable lead-acid cells (or sometimes Nickel-Iron Edison cells). In the absence of railroad-generated power, a limited number of commercial ac power drops were installed at central locations along each railroad line, and the railroads then used their own poles and wires to distribute the power. The commercial ac power would be converted to 550 V ac (or higher), for local distribution along the railroad. This was often done using a single pair of #6 wires carried on the same poles as the signal system linewires running parallel to the tracks. A local pole-mounted transformer at each signal equipment location would step the voltage back down to 110 V ac for use by the battery chargers and rectifiers. Drawing only a very light load at each signal location, an ABS system covering up to 50 miles of track could be serviced by a single pair of wires in this fashion. 4-69

Overview of Railroad Signal Circuits

Primary Batteries The primary batteries used for most early dc track circuits used caustic soda type primary cells (known as Lalande cells), which had a zinc anode, a copper (actually cupric oxide) cathode, and a sodium hydroxide electrolyte. These had an open-circuit voltage of approximately 0.95 Volts, but this dropped to about 0.6 volts under heavy loads. New cells were usually assembled on-site from individual components. The powdered electrolyte was mixed with water in the large (> 1 gallon) glass or ceramic jar that served as the battery case. The anode and cathode would be attached to binding posts protruding through the removable lid of the jar, to be immersed in the electrolyte solution when the lid was installed. After installing the lid, a thin layer of battery oil would then be added through a small funnel-shaped hole in the lid, to reduce evaporation of the water in the electrolyte.

Figure 4-56 Primary Railroad Signal Batteries (Lalande Cells). New (Left), and Discharged (Right)

Battery maintenance consisted of replacing any water lost from the electrolyte through evaporation, and rebuilding each cell when it had been exhausted. Railroad personnel known as Battery Maintainers always carried water, new electrodes, powdered caustic soda (sodium hydroxide), battery oil, and the other small parts needed to rebuild a cell. When a cell became exhausted, the liquid contents were customarily dumped along the railroad right-of-way, the old electrodes discarded trackside, and a fresh batch of electrolyte was mixed up in the same jar. New electrodes were installed onto the underside of the lid, and the cell was returned to service. These disposal practices obviously pre-dated hazardous material disposal laws, and have since been discontinued. However, some lingering effects from this, usually in the form of unusual soil

4-70

Overview of Railroad Signal Circuits

conductivity in the area of old signal equipment enclosures and their nearby battery boxes can still be seen. But not all primary batteries were of the Lalande caustic-soda type. Air-depolarized carbon-zinc primary batteries were once used, and continue to be found in dwindling numbers on the most remote sections of railways in the western United States. Rechargeable Batteries Over time, the cost of maintaining and rebuilding primary batteries, along with the expanding availability and reliability of commercial ac power, caused railroads to switch to rechargeable lead-acid or nickel-cadmium batteries, used in conjunction with rectifier/charger circuits. The lengthening of track circuits has also enabled the elimination of batteries at many ABS locations or cut-sections where reliable ac power is unobtainable at reasonable cost. Using this method, the batteries are only discharged during ac power outages, or periods of unusual load, and can have very long service lifetimes, if carefully maintained.

Figure 4-57 Modern Single Lead-Acid (or Ni-Cd) Battery Cell

Cells are generally available in sizes from around 100 Amp-Hours to over 1000 Amp-Hours, with cells in the 120 to 400 Amp-Hour range being the most common. The size of the battery bank constructed is determined by the size of the load, the voltage required, and the number of hours of stand-by time required in the event of an ac power failure. The required stand-by time is usually specified by each states public utilities commission (P.U.C.), or may be specified by the railroad based on past experience. Twelve hours is a common requirement, but some localities require much more than this.

4-71

Overview of Railroad Signal Circuits

Transmitters The basic dc track circuit transmitter uses a single-cell lead-acid rechargeable battery, and an adjustable wire-wound resistor of a few ohms. As this circuit can present an excessive load (rail-to-rail shunt) for other track circuits using ac currents on the tracks, a battery choke is often installed in series with the battery and resistor, in order to minimize this effect.

Figure 4-58 Simple Schematic Drawing of Cell, Resistor, and Battery Choke

A battery choke consists of an iron-core inductor normally wound on a gapped C core. This inductor is high enough in inductance to provide several ohms of reactance at the lowest ac signaling frequency used on the tracks, and is physically large enough to avoid significant core saturation at the dc current levels usually employed by the track circuit.

Figure 4-59 Drawing of Typical Track Circuit Battery Choke (Safetran Systems Corp.)

4-72

Overview of Railroad Signal Circuits

Coded dc track circuits have a transmitter that transmits a continuous stream of dc pulses down the track. The duty cycle is roughly 50%, and this type of track circuit not only determines the occupancy of the track circuit, it can also transmit a limited amount of information from one end of the circuit to the other by changing the frequency of the pulses. The different pulse frequencies are generated using several finely built electromechanical (pendulum-type) oscillators. These oscillators are equipped with relay-type contacts, which make and break with each full cycle of the pendulum. The frequency transmitted is determined by selecting which electromechanical (pendulum-type) oscillator is used to generate the dc pulses. Normally, coding rates of 75, 120, or 180 pulses-per-minute are used, but there are also others. Thus, a coded dc track circuit transmitter looks essentially like a battery that is connected to the tracks by a set of relay contacts that open and close to generate dc pulses. However, additional filtering equipment is often added between the coding relay and the track, in order to reduce the harmonic content of the pulses. This is normally the case when there are Motion Sensors or Crossing Predictors connected to the tracks within the same coded dc track circuit. This concept of applying dc pulses to the track can easily be extended to include ac signals as well. By substituting a source of ac energy of a particular frequency for the battery, a coded ac track circuit transmitter can be created using the same coding relay. Decoding this type of signal is more complicated, but the transmitter is very similar to that of the coded dc track circuit. Another twist on the idea of using a coded dc track circuit not only determines track occupancy, but also transmits a limited amount of information down the tracks by using transmitter electronics that generate widely-spaced dc pulse pairs. These pulse pairs carry information not in the rate (pulse frequency) at which they are generated, but rather in the amount of time between the pulses in the pair. This can be used in place of open linewires or cable to convey track occupancy information. Transmitters in most ac track circuits use a transformer to couple the ac signal onto the track, and may also use a coupling capacitor to block any dc voltage present on the tracks.

Figure 4-60 Transmitter Coupling Transformer and Capacitor

4-73

Overview of Railroad Signal Circuits

This is the case for Motion Sensors and Crossing Predictors, whose electronics are both capacitor-coupled, and transformer-coupled to the tracks. There are also other types of track circuits whose transmitters apply either ac or dc voltages to the track. These will each be discussed separately, as each specific type of equipment is introduced. Receivers Correct Reception is the Key to Safety At the opposite end of a track circuit from the energy source (the transmitter or track battery) is the track circuits receiver. Robinsons original 1872 track circuit patent describes a circuit that uses the coil of a relay (the track relay) to detect the energy supplied by the battery connected to the rails at the opposite end of the track circuit. In effect, the relay acts as the receiver for a dc track circuit.

Figure 4-61 Typical Instrument Case Containing Shelf Mounted Vital Relays

Although this is still a very common type of track circuit, many modern track circuits have replaced relay coils with electronics. This is especially true when we expand our discussion to include those that do not use steady dc currents supplied more-or-less directly from a battery/rectifier. Track circuits have been constructed which use ac current (40 Hz, 60 Hz, 100 Hz, or other frequencies), pulsed-dc current, and even ac current with specific Amplitude Modulation or Frequency Modulation impressed upon it. The receivers used to detect these types of electrical 4-74

Overview of Railroad Signal Circuits

signals employ tuned filters, detector/demodulator sections, threshold comparators, and even Digital Signal Processing techniques. But, in all cases, the specific design of the energy-detection device (the receiver) is driven by the characteristics of the electrical energy placed on the tracks by the track circuits transmitter.

Figure 4-62 Typical Instrument Case Containing Plug-in Vital Relays

In order to discuss the characteristics of many railroad signal system components, we will need to introduce a few terms here. We will refine these concepts later, as we discuss how they apply to each particular piece of signaling equipment. The first of these terms is: vital. Simply put, a vital circuit is one designed to be fail-safe. Vitality, in the railroad sense, always has something to do with safety. If a circuit is vital to safety, then its referred to as a vital circuit. Or to put it a bit more morbidly, if a vital circuit fails to operate correctly, someone may die. The FRA doesnt necessarily use the term vital, preferring instead to write things like rule 236.5: 236.5 Design of control circuits on closed circuit principle. All control circuits the functioning of which affects safety of train operation shall be designed on the closed circuit principle, except circuits for roadway equipment of intermittent automatic train stop system. With this rule we can extend our definition of vital to include the term: closed circuit principle. This concept is also called the normally-energized principle. This means that the circuit is normally energized, and any failure (including the power supply) will cause the circuit to put the system in its safest possible state. For example, the gate mechanisms used at grade crossings are designed so that they require a very small current at all times to keep the gates vertical. Should this current be interrupted, a 4-75

Overview of Railroad Signal Circuits

small ratcheting latch assembly known as a hold clear will release, and the gate will descend slowly to its horizontal position by gravity alone. Track circuit receivers, whether they are simple relays, or something more complex, must also be designed this way. Since much of the safety and fail-safe characteristics of a track circuit depend on correctly interpreting the electrical signals received from the track, it should come as no surprise that a great deal of effort goes into the design of the receiver. A track circuit receiver must never misinterpret the electrical signals that they get from the track in such a way as to cause the railroad signal system to permit an unsafe condition. Railroad track relays, the simplest of track circuit receivers, have many design features included specifically to prevent certain failure modes from occurring. For example, the contacts on most railroad relays (and all relays used for vital functions) are constructed of dissimilar materials, in order to prevent welding of the contacts should an excessive current be passed through them. One contact is metallic, and usually made of silver, while the other is usually made of silverimpregnated carbon. During an over-current condition, the contacts may experience extreme heating, but no fusion can occur between the metallic and non-metallic contacts. This particular contact design makes sure that when a relay is de-energized, its armature will always be able to move to the de-energized position. Furthermore, although some railroad vital relays have springs in them to increase their speed of operation, all vital relays have armatures that are designed to drop-away to their de-energized position by gravity alone. Even though there is a wide variety of track circuit types, and hence, a wide variety of track circuit receiver designs, one thing remains nearly constant for all track circuits. That is, the occupancy of the track is almost always determined from the amplitude of the received signal. More complicated receivers, such as those used in ac, pulsed-dc, Audio-Frequency Overlay, Motion Sensor, or Predictor circuits, have many features designed to ensure proper operation. Generally, these take one or more of the following forms: 1. The receiver should only accept (and respond permissively to) an input signal of the correct frequency, phase, amplitude, pulse-width, timing, modulation, code, polarity, etc. 2. The receiver should not respond unsafely to the presence of any other electrical signals coming from the track. 3. If conditions are detected which could prevent the receiver from being able to correctly receive and process the signal that the receiver was designed to process, then the system must fail-safe and take action to protect the safety of train operations and the safety of the public. 4. In the case of electronic track circuit receivers, constructed using components which are not intrinsically vital in and of themselves, the receiver must be capable of testing itself during normal operations, and any internal failures detected must initiate a fail-safe action, as in item 3, above. 5. Self-testing of the receiver must occur often enough to allow for detection of internal failures before that particular failure could appreciably affect the overall safety of the system. 4-76

Overview of Railroad Signal Circuits

A functioning signal system normally complies with the five requirements above. When it doesnt, this is called a wrong-side failure. A wrong-side failure is said to have occurred whenever a railroad signal system, or system component, fails to operate correctly and that failure does not initiate a fail-safe action. In contrast, a right-side failure is said to have occurred whenever a failure was automatically detected, and the signal system or component initiated a fail-safe action. Any device which exhibits wrong-side failure modes cannot be considered vital, with the following exceptions. One of the underlying principles of the railroad signal system is that it is single-fault tolerant. That is, any single failure which could affect the safety of the system must initiate a fail-safe action. However, the signal system is not required to be able to handle multiple failures that occur simultaneously. Most railroad signal equipment also has some degree of multiple-fault tolerance, even though it is not strictly required by FRA rules. In simple dc track circuits, the receiver (a track relay) achieves its vitality by having as little complexity as possible, and by incorporating specific measures to prevent all known sources of wrong-side failures. Most other track circuit receivers rely on other methods to achieve vitality such as the use of multiple independent signal processing channels. The outputs of the channels are continually compared to confirm agreement. When the outputs differ significantly, then a failure has been detected, and the system must take fail-safe action. In general, fail-safe action consists of declaring the presence of a train even when there isnt one. This results in false activations of grade crossing warning systems, and wayside signals falsely indicating STOP for the train traffic. These right-side failures result in lost time, lost revenue, and increased maintenance costs, but are far preferable to the wrong-side failures which can cost human lives and millions of dollars. The ideal track circuit receiver would have no wrong-side failure modes, and would never initiate a right-side fail-safe action unnecessarily. Unfortunately, external interference such as that from ac power lines can sometimes rise to levels that initiate right-side failures. Signal equipment manufacturers are continuously trying to make their equipment more immune to this interference, but this concern must always be balanced with others, including human safety, compatibility with other existing equipment, and economics. Our purpose in this handbook is to help the reader understand the current realities of ac interference as they apply to railroad signaling, and provide the means to improve the situation. The Output of the Track Circuit Now that we have looked at the source of energy (transmitter), transmission line (track), and the receiver of a track circuit, lets examine the output of the track circuit. In the case of a track relay, the output is integrated into the receiver. The coil of the relay acts to convert the received current into magnetic flux, drawing the armature of the relay towards the coil. This opens a contact or contacts, and closes others. The output of a relay is just the relay contacts themselves, which when connected to a source of energy such as a battery can control a

4-77

Overview of Railroad Signal Circuits

wide variety of other devices and circuits including other relays, crossing warning devices, track switches, and wayside signal lights. Ultimately our analysis of a track relay as a track circuit receiver ends in a very simple conclusion. The track relay is either energized (up) or de-energized (down). These up and down descriptions come from the positions of the gravity-actuated armatures of railroad relays. Therefore the output of a relay is essentially binary, containing a single bit of information. A track relays basic function is to determine whether or not the track is occupied. This is done with a simple threshold determination. If the received signal level is above the threshold, the track circuit is considered to be unoccupied. If it is below the threshold, the track circuit is occupied (or possibly damaged). More sophisticated receivers of track circuit energy, such as those in Motion Sensors, Crossing Predictors, and some audio-frequency track circuits, can do a much more detailed analysis of the received signal. Aside from the usual filtering, amplification, detection, and other operations that extract the desired signal from the mixture of electrical signals on the track, these devices can do some extensive signal processing. The various characteristics of the signal received from the transmitter, such as amplitude, frequency, phase, modulation, etc. may be individually measured. These discrete measurements can then be combined mathematically to determine what the state of the output should be. (This will determine whether or not the crossings warning devices should be activated.) The actual output of a Motion Sensor, Crossing Predictor, or audio-frequency track circuit will generally be in the form of a vital relay drive output. That is, the output will consist of 12 Volts dc, with enough current available to drive the coil (usually 500 Ohms) of a standard signaling relay. This output is normally energized when the track circuit is un-occupied, just as the coil of an ordinary track relay is. Even though these devices are more sophisticated, and may have more than one output, the essential nature of each output remains the same. Each track circuit output provides the signal system with just one bit of information. For audio-frequency track circuits, the output indicates track occupancy (and integrity), just as the output of a relay does. For Motion Sensors and Predictors, the output indicates whether the crossing warning devices should be activated or not. And like all other parts of a vital signal system, these outputs must also be vital. The presence of ac interference is of greater concern where electronic relay-drive outputs are used, as the diodes in these circuits can potentially rectify induced ac and falsely energize the relay coil connected to the relay-drive output. However, the electromagnetic field strengths necessary to produce this effect are extremely rare. Shunting In Robinsons track circuit, occupancy of the track by a train results in a short-circuiting of the energy being fed to the Receiver (Track Relay). This short-circuiting action is termed shunting. Robinsons track circuit was revolutionary, because shunting of the track circuit by

4-78

Overview of Railroad Signal Circuits

each axle of the train was essentially automatic when the train entered the circuit, and generally quite reliable.

Figure 4-63 Shunting of the Track Circuit

The degree to which a train or shunt is consistently able to shunt current away from the track circuit receiver is called shunting performance. Shunting performance is never black-andwhite, and always a matter of percentages. In any real-world situation the shunt presented by a train never completely deprives the track relay of current. Under ordinary circumstances, the residual current flowing through the coil of the track relay while a track is shunted is a very small percentage of the track relays drop-away value14.

14

The drop-away value is the maximum current at which the armature is guaranteed to move to the de-energized position.

4-79

Overview of Railroad Signal Circuits

Figure 4-64 Shunting Efficiency of Track Circuit

Given unfavorable conditions, however, a poorly shunting train may fail to reduce the level of 15 received current below the value necessary to open the track relays front contacts , creating a dangerous false-proceed condition (also known as a false clear). Shunting difficulties are usually observed in areas where either the rate of railhead wear is unusually low, or the rate of contamination or corrosion is unusually high. Despite the very high contact pressures at the wheel-rail interface, these factors can lead to the build-up of an insulating film that is not readily punctured. This effect can be readily seen on electrified railways when trains coast without drawing significant amounts of propulsion current. In severe cases, the track relay will be observed to pick-up and then drop again as traction power is removed and applied by the train operator. In this case, the propulsion current returned through the wheels and into the rails is acting as a wetting or whetting current (see below).

15

The terms front and back are used to refer to the contacts of the relay which are touched by the armature contacts in the energized and de-energized positions, respectively. In this context, energized, up, and front are essentially equivalent. (Think: energized/up/front, de-energized/down/back.) The armature contacts are referred to as the heel contacts.

4-80

Overview of Railroad Signal Circuits

But these same symptoms can also be induced on diesel-electric railroads by purely mechanical means. For example: under a heavy tractive effort (e.g. accelerating away from a stop), the shear forces present in the contact patches between the wheels and rails can cause a small amount of slip. Often this is all that is needed to rupture the insulating film on the railheads, which allows the axles of the train to shunt properly. But when the throttle is retarded, allowing the train to coast, the wheels may begin to ride up onto the film without rupturing it. This can prevent the wheels and axles of the train from shunting properly, and can cause the track relay to pick up even though the track circuit is actually occupied. The contributing factors to a case of poor shunting can vary widely, but many of them are well known. Some of these include: Sand, rust, leaves, spilled loads, or other insulating contaminants: These can form insulating films on the railhead. The use of track circuits having low rail-rail voltages: This minimizes the voltage available to punch through the insulating or semi-conductive films on the railheads. Light axle loads, usually less than 8-12 tons per axle: This can help minimize the wear rate at the wheel/rail interface, and therefore minimizes the amount of clean, conductive steel exposed on both the rails and wheels. Short trains having fewer axles: This factor has a dual effect, in that it both minimizes the wear rate at the wheel/rail interface, but also reduces the number of axles placed in parallel across the track circuit. If conditions prevent an axle in the track circuit from providing a 0.06-Ohm shunt or less, then the shunting performance of the train will rely on other axles in the track circuit to provide the needed shunt. Infrequent train traffic: This reduces the wear rate on the railheads, and can allow a complex mixture of contaminants to accumulate. However, rust is still usually a primary ingredient in the mix. The use of modern composite brake shoes (which replaced the earlier soft-iron brake shoes): Composite brake shoes use a binding agent that can sometimes coat the tread of the wheels. The older soft-iron brake shoes tended to clean most contaminants from the tread of the wheels. Better maintenance of wheel-rail profiles: This tends to reduce the amount of wear at the wheel/rail interface. Local environmental conditions which are unusually corrosive: These increase the corrosion (rusting) rate of the steel rails. This is often a problem in coastal areas, but can also result from local industrial emissions. Radial steering trucks on the trains: These reduce the mechanically-undesirable flange contact with gage faces of the rail on tangent (straight) track, which reduces friction. Unfortunately, it is this flanging contact that has historically been a component of the shunting performance of trains in track circuits. Absence of propulsion return current through the track: This acts as a wetting current (see above).

4-81

Overview of Railroad Signal Circuits

The presence of grease on the railhead or wheels: Automatically applied rail grease, often used to lubricate the flanges of the wheels, before or as they pass around sharp curves, does 16 not by itself appear to be a problem . However, it may act as a binding agent for other substances that have a more adverse effect on shunting. Limitations inherent to the type of track circuit.

It is important to mention poor shunting in a discussion of EMC issues involving ac power lines and railroad signal circuits. This is because resolving an ac interference problem that formerly raised the rail-to-rail voltages on the track may reveal an underlying problem with shunting. Remedies for Poor Shunting The vast majority of track circuits function properly with virtually no attention or adjustment required after their initial set-up during installation. Some track circuits, however, experience chronic shunting problems. This is rarely the fault of the track circuit hardware itself, and is usually is much more a result of the conditions on the tracks. On low-speed or seldom used trackage, where rusty or greasy rail creates a chronic poorshunting problem, stainless steel weld beads are sometimes applied to the top of the railhead. The use of thin copper strips pressed into shallow longitudinal groves cut into the railhead has also proven successful in hump yards17. These techniques cannot be applied to mainline trackage due to their cost and adverse effect on the rail life, especially in heavy-tonnage territory. Another technique for improving shunting performance was borrowed from the telecommunications field. The telephone industry recognized early-on that electrical circuits carrying very small signals were often unreliable. This was due to the effects of localized resistance on the mating faces of contacts and the spurious rectification that could occur at junctions of dissimilar metals or their oxides. This difficulty was overcome by adding a small dc wetting current. In this way, the low-level ac signal carrying the audio information (a telephone conversation) is superimposed on a steady dc signal used for control purposes (onhook/off-hook). This dc signal therefore provides all of the power necessary to operate the subscribers telephone instrument, and greatly reduces line noise in exchange for the burden of only a few milliamps per subscriber line Today, the wetting-current concept has been applied to a variety of track circuits where satisfactory shunting has been hard to achieve. Differential (rail-to-rail) voltages caused by induction from utility lines may, in some cases, actually have a beneficial effect in providing a wetting current which improves shunting performance.

16

The Association of American Railroads conducted a study in the 1990s which showed no significant positive correlation between the use of rail lubricators and the occurrence of shunting problems within the island track circuits at grade crossings with a history of shunting problems. A hump yard is a type of railroad switching yard that is built on a slope. Cars are moved to the top of the slope (hump) and released, after which they are allowed to roll downhill. Remotely-controlled track switches and special braking mechanisms control the movement of the car as it is routed towards the end of the train to which it will be added.

17

4-82

Overview of Railroad Signal Circuits

Another version of this idea is the Track Circuit Assister (TCA), which has been used in the UK since the early 1990s to counter the effects of poor shunting performance. Generating a roughly 165 KHz signal, the TCA loop mounted between axles on each truck induces a substantial railrail voltage beneath the train. TCA current circulating through the wheel sets provides a wetting action that breaks down the insulating wheel-rail film. At 165 KHz, the TCA signal is usually far above the frequency range of any track circuit and therefore does not interfere with most wayside track circuit equipment. Decomposing crushed leaves can form a tough crust on top of the rails that cannot be brokendown by the TCA. For areas where leaves are a seasonal problem, Track Circuit Assister Interference Detector (TCAID) has been installed. As the name implies, TCAID is a wayside element that senses the presence of the TCA signal and independently shunts the track circuit, forcing an indication of track occupancy. TCAID receivers are typically installed at 200-meter intervals along the track. The primary application of TCAID is in situations where the railroad is unable to remove deciduous trees that are in close proximity to the track. An earlier and somewhat simpler implementation of TCAID is the use of treadle wheel detectors to knock-down the track circuit as a train enters the circuit. This is essentially a mechanical means of detecting the presence of the train, which then shunts the track circuit for a period of time. This scheme is widely used in Europe. Combined with the proper time delay, this scheme is highly effective at preventing loss-of-shunt episodes. The development of the TCA follows early efforts in the U.S.A. during the 1950s to improve the shunting performance of lightweight trains using a similar on-board signal injection technique. Following a serious accident in Palmer, Massachusetts, which was attributed to a stopped train resting on rails whose heads were coated with a layer of sand, a number of cars were fitted with axle-mounted transformers that induced a circulating current into the wheelsets and the rail beneath them. This scheme proved only partially effective. Railroad Track Circuits One at a Time Much of the history of railroad signaling is still with us, in the form of older equipment that is still in operation. In this section, well take a look at each type of equipment still commonly in use in greater detail. Steady Energy DC Track Circuits Early dc track circuits employed DC Neutral Vital Relays with coil resistances in the 1- to 25ohm range. Copper-zinc wet-cell primary batteries with open circuit voltages of 0.6-0.95 Volts were usually employed, providing roughly 0.8 Volts under load at the beginning of their lifetime. These cells could be hooked up directly to the rails, since their high internal resistance prevented excessive current flow while the train occupied the track circuit. Later high voltage cells provided open circuit voltage of up to 1.5 Vdc.

4-83

Overview of Railroad Signal Circuits

Figure 4-65 Elementary Steady Energy DC Track Circuit

Early track relays had relatively poor and unstable operating characteristics, thus the maximum operable length of a track circuit was limited to somewhere in the 3,000 to 5,000 range. A major problem for the relays was poor Release Ratio, which is the ratio between the track relays drop-away and pick-up current values. The drop-away current was typically only 30-40% of the pick-up value. This required a large change in the coil current (caused by the train on the track) in order to de-energize the relay. Because these relays employed soft-iron magnetic cores, pole pieces, and armatures, the residual magnetism present increased over time, which further degraded the release ratio. Old relays might have a release ratio of 25% or less. As this release ratio declined over time, a standard (0.06-ohms) test shunt placed across the track circuit might no longer be sufficient to release the track relay. Under these circumstances, a false proceed condition will exist if the track relay fails to drop when the corresponding section of track is occupied by a train. For this reason, the FRA requires that relays with soft-iron magnetic structures have their operating characteristics tested at least once every two years (FRA Rule 236.106c). For this reason, most railroads have phased these soft-iron relays out at every opportunity. This testing is necessary to detect diminished drop-away values caused by magnetic aging and other defects such as worn stop pins. Relays exhibiting low drop-away values due to residual magnetism must have their cores re-annealed if they are to be returned to service. Early dc track relays of this type typically required shop overhaul after only two to four years of service. The magnetic aging problem of the soft-iron cores was largely eliminated with the adoption of 3-4% silicon steel for magnetic circuit materials beginning in the mid-1920s. This has resulted in a magnetically harder structure. Modern dc track relays have release ratios in the range of 65-85%, and require testing only once every four years. 4-84

Overview of Railroad Signal Circuits

Unfortunately, the permeability of silicon steel is somewhat less than that of soft iron, resulting in a slight sacrifice in other aspects of the track relays performance. This handicap has been compensated for by a rigorous mathematical analysis of the track circuit parameters, better rail bonding, more stable feed-end voltages, and various improvements in relay design. Additionally, the dc track relays used today are usually magnetically biased (Biased-Neutral DC Relays), so that dc current of the wrong polarity cannot falsely pick-up the relay. This improves the degree of protection from multiple failed insulated joints, which could energize the relays in two adjacent track circuits if both joints separating them failed. With the exception of alternating the battery polarity on each successive track circuit, the rest of the circuitry remains the same with this type of relay. The dc track circuit continues to be widely used for short-to-medium range track circuits in nonelectrified (no electric trains) territory. Track circuit lengths up to 6000 feet are feasible without special consideration. Provided that no foreign dc current is present, maximum operable lengths of up to 18,000 feet are possible using modern 0.5-ohm track relays having high release ratios. However, because other more-reliable options now exist, simple dc track circuits longer than 6000 feet are generally not recommended. Another way to construct a dc track circuit is to connect the battery to the rails in the center of the track circuit, and place a relay at each end. The contacts of the relays are then wired in series with one another to produce the occupancy indication. Center-fed dc track circuits up to 15,000 feet in length have been widely used. Center-fed track circuits have several handicaps, including an inherent difficulty in detecting broken rails near the center of the circuit. With the advent of electronic track circuits offering bi-directional communications capability via the track, new installations of the center-fed dc track circuit have all but ceased.

Figure 4-66 Center-Fed DC Track Circuit

4-85

Overview of Railroad Signal Circuits

Polarized DC Track Circuits A significant variant of the steady energy dc track circuit is the polarized dc track circuit. Early polarized dc track circuits used polarity-responsive dc track relays having two armatures. One armature operated contacts that closed whenever current of either polarity flowed through the coils. The other armature was responsive to polarity (usually by means of permanent magnets), and its contacts moved in one direction or the other depending upon the polarity of the energy applied to the coils. In a simple 3-aspect signal system (one using red, yellow, and green wayside signals), the aspect displayed depends not only on the occupancy of the block ahead, but also on the occupancy of the block beyond it. This means that the occupancy information from a distant block must be transmitted at least the length of one block to the logic circuits controlling the wayside signal. The most common signal practice used to be to transmit this information via the linewires running parallel to the tracks. These are the often-misidentified telegraph lines strung on wooden poles next to the tracks. Later, these open linewires were replaced by overhead cable, underground cable, or even vital radio-based equipment. The advantage of using a polarized dc track circuit is that it can eliminate the need for at least some of these linewires or cables. By encoding the occupancy information of a block into the polarity of the dc voltages applied to the rails of the blocks on either side of an occupied block, a determination can be made regarding the correct color for the signal lights on the adjacent blocks. Heres how it works:
Table 4-1 Signal Aspect (Color) Displayed for Polarized DC Track Circuits Polarity of Track Voltage Received Positive Negative No Voltage (track shunted) Meaning The next block ahead is unoccupied, and the block beyond it is also unoccupied The next block ahead is unoccupied, but the block beyond it is occupied The next block ahead is occupied Signal Color Displayed Green Yellow Red

The key to the transfer of information is that a wayside signal which receives no voltage (occupied block) on one side of the insulated joints will be Red for trains approaching that block. Furthermore, this signal will change the polarity of the track circuit voltage transmitted towards the preceding signal, so that the preceding signal will turn yellow. In this way, signal control information is transmitted via the rails, which still simultaneously serves as the means of train detection. Later, polarized track circuits employed back-to-back biased-neutral track relays. Because the polarized dc track circuit has several limitations, its usage is dwindling. Chief among these is the additional equipment required to provide signals for trains running in either direction 4-86

Overview of Railroad Signal Circuits

(bi-directional signaling). Another limitation is an inherent difficulty in detecting multiple insulated joint failures.

Figure 4-67 Polarized DC Track Circuits for Transmission of Signal Control Information via the Rails

A later improvement to the polarized dc track relay was the addition of retaining coils (Retained-Neutral Polarized Relay), permitting the neutral front contacts to remain closed during a polarity change. This feature prevented colorlight wayside signals from momentarily turning red during a yellow-to-green transition. Coded DC Track Circuits Introduced in the 1930s, the coded dc track circuit offered many advantages over the simpler steady-energy dc type and other track circuits that were available at the time. In a coded track circuit, voltage is intermittently applied at the feed-end of the circuit. The code rate is typically 75, 120 or 180 pulses-per-minute, and the duty cycle is nominally 50%. At the receive end of the circuit, a code-following track relay repeats the code pattern. Using slow-release relays or tuned L-C circuit decoders, the continuous transmission of code is detected and translated. This action maintains closure of the front contacts of a track repeater relay, which behaves the same as the track relay for a steady-energy dc track circuit.

4-87

Overview of Railroad Signal Circuits

As with any other track circuit, the presence of a shunt anywhere within the track circuit reduces the current received by the track relay, causing the track to be declared occupied. Coded dc track circuits typically have a maximum operable length in the 7,000 to 15,000 range, although signal equipment manufacturers are always trying to improve upon this.

Figure 4-68 Elementary Coded DC Track Circuit

Foremost among the advantages of the coded track circuit is that the release ratio of the track relay effectively approaches 100%. This is an ideal yet unobtainable value when using ordinary neutral dc track relays. So how is this possible? As the code-following track relay is continuously cycling between its energized, and deenergized positions when the track is unoccupied, the trains shunt need only reduce the amount of current flowing through the relays coils to something less than the pick-up value of the track relay. The shunt does not need to reduce the residual current to less than the code-following track relays minimum drop-away current as it does with a convention track circuit. This feature permits much longer track circuits, improved shunting sensitivity, and greater assurance of broken-rail detection. Using modulated or coded energy also greatly improves the security of the track circuit, as there is a danger of falsely energizing ordinary dc track relays in areas where foreign dc currents may be present. Typical sources of foreign dc energy include electrified railroads using dc-propulsion current, cathodic pipeline protection systems, high-voltage dc transmission lines, geomagnetic currents, industrial plants, mines, galvanic action due to ballast contamination, and many other similar sources. In the case of coded dc track circuits, should the code-following track relay be held steady in its energized position by a foreign current, this will still cause the decoding circuit to turn off its output, which de-energizes the track repeater relay in the same manner as if the track circuit were 4-88

Overview of Railroad Signal Circuits

occupied by a train. Although coded dc track circuits are sensitive to foreign dc energy, the coded circuit will often prove workable where steady energy circuits do not function satisfactorily. With the advent of coded dc track circuits, new investigations into track circuit parameters uncovered the phenomenon of track storage effect (TSE). Previously, it had been observed in ordinary dc track circuits that under the right track conditions, the track relay would remain in its energized position for several seconds after the feed-end energy was removed. Analysis showed that wet and dirty track ballast tended to behave as both a polarized capacitor and a galvanic cell, with the rails serving as the plates of both. Together, these actions mimic the behavior of a weak storage cell. Contaminated ballast, and ballast materials being in contact with the rails are the leading contributors to TSE. For coded track circuits, the battery action of the track itself is an important factor that limits the maximum operable track circuit length. Regardless of the type of track circuit used, railroad tracks are generally inefficient transmission lines. Because coded track circuits operate over very long lengths of track, the sensitivity of the code-following track relay used must be very high. As track circuits work primarily in the current domain, as opposed to the voltage domain, this means that the resistance of the track relays coil must be minimized. The coil of a typical code-following track relay has a resistance of only 0.3 ohms and a pick-up current of 410 milliamperes, the most sensitive of any class of track circuit. Because the track relays coil has such a low inductance and resistance, being made of as few as 13 turns, the windings are sometimes susceptible to damage caused by lightning strikes, and are even more susceptible to surge events from ac power lines. And because the armature has a very small mass, a code following track relay can usually respond to 60 Hz ac interference. Some track relays have resonating windings to suppress 60 Hz interference. It is this multiplicity of expensive, sensitive, and continuously moving parts that has prevented the coded dc track circuit from achieving dominance over the simple, low-cost, steady energy dc track circuit. By 1945, the electromechanically-coded dc track circuit had evolved into a complex, bidirectional signal system that eliminated the need for line wires. Although new installations of this type ceased by the early 1980s, thousands of miles of rate-coded dc track remain in service today throughout North America. Because of the harmonic content present in the pulses from the code-transmitting relay, appropriate filters must often be used to prevent interference with Motion Sensors and Crossing Predictors installed on the same tracks as coded dc track circuits. The effect of these filters is to low-pass the pulses and round off their corners. Alternatively, sometimes the operating frequencies of the Motion Sensors and Crossing Predictors are changed in an attempt to move away from the bulk of the harmonic content. Trakode The cost and complexity of rate-coded dc track circuits prompted the development of so-called slow-code track circuit in the early 1950s. The first and most widely used type

4-89

Overview of Railroad Signal Circuits

was manufactured by the General Railway Signal Corporation (G.R.S., now a part of Alstom Signaling Inc.), and was called Trakode. This form of track circuit utilizes brief pulses of about 250 milliseconds duration transmitted at roughly two-second intervals (either 33 or 29 pulses per minute). Decoding circuitry is relatively simple, inexpensive and not dependent on the variable timing characteristics of relays. By inverting the polarity of impulses, or transmitting pulse pairs of opposite polarity, signal control intelligence may be transmitted via the tracks, thus eliminating the need for line wires. The slow code rate and alternating polarity of Trakode pulses helps to overcome the track storage effect, permitting substantially longer circuits. One circuit was reportedly stretched to 35,000 feet during winter months when the frozen ground created unusually high ballast resistance conditions. However, the slow code rates used by this system had a drawback, in that detection of a shunt (train on the track circuit) could take as long as three seconds. Trakode was widely installed throughout the 1960s and 1970s. Recurring difficulty with burned or welded code pulse transmitter contacts and susceptibility to 60-Hz ac interference has led to its replacement by electronic track circuits in many areas. As with other coded dc track circuits, the varying impedance presented to the track by the transmitter circuit requires the addition of isolation inductors or filters where Crossing Predictors or Motion Sensors are used. Trakode II, a microprocessor-based system that was compatible with the earlier relay-based systems, was introduced by G.R.S. in 1982, but failed to win a substantial market share. PMTC (Pulse-Modulated Track Circuit) The era of electronic track circuits began with introduction of the Pulse-Modulated Track Circuit, or PMTC, by Marquardt corporation (later part of Safetran Systems Corporation) in the mid-1960s. In larger installations of this type, train detection is done using ordinary center-fed dc track circuits. The PMTC superimposes audio-frequency tone bursts that transmit signal control information (on top of the dc energy) down the rails. Because the underlying dc track circuit provides the needed shunt and broken rail detection, the overlaid audio frequency signals can be transmitted and decoded without complex (i.e. vital) circuitry. Track circuit lengths up to 15,000 feet have been successfully used. Substantial amounts of PMTC equipment is still in service. However, this system is now considered obsolete, and has been out of production for some time. As one of the earlier audio-frequency track circuits, it was more susceptible to ac interference than some later track circuits.

4-90

Overview of Railroad Signal Circuits

Figure 4-69 Typical Electronic Track Circuit

Electro Code Electro Code is presently the most widely used family of electronic track circuits used in North America, and has become a de-facto standard. Its originated with an effort in the late 1970s by a Midwestern railroad to replace troublesome components in their rate-coded dc track circuits. The timing of the dc pulse pairs used by Electro Code closely resemble the 75, 120, 180, and 270 pulse-per-minute code rates found in earlier coded dc track circuits. As with its predecessors, Electro Code applies a series of single-polarity dc pulses to the track. Due to track storage effect, the 112-millisecond pulse can be significantly distorted, and the maximum operable track circuit length is roughly 15,000 feet. The amount of information that can be encoded into the pulses transmitted along the rails is significantly higher than that of earlier systems. The messages transmitted (the state of the corresponding outputs) are a combination of both vital and non-vital bits. In its standard form, Electro Code can control up to five signal aspects, and one non-vital auxiliary function, each of which consists of one bit of information. Since the Electro Code receiver uses pulses of only one polarity, and its receiver is polarity-sensitive, the polarity is usually alternated at every set of insulated joints. This helps prevent wrong-side failures as a result of a pair of failed insulated joints. The Electro Code II system was the first complete, self-contained Automatic Block Signaling (ABS) system in-a-box. All of the inputs, outputs, and logic functions needed to control bidirectional ABS signals, are contained in one chassis. All electrical power to the chassis is provided from a single 12-Volt dc source. Later versions of Electro Code have also eliminated the need for external vital relays for signal lighting. At roughly $500 apiece, with several such vital relays at each signal location, the railroads view this cost savings as significant. 4-91

Overview of Railroad Signal Circuits

The basic Electro Code track circuit has now been integrated into solid-state processor interlocking systems such as Microlok II and Electro Code 5. Later versions also included additional built-in filtering, which significantly increased the level of ac interference which the system could withstand.

Figure 4-70 Pulse Timing of Electro Code II, 3, and 4

Electrified Electro Code In order to extend the use of Electro Code to electrified railroads, an add-on unit was developed which was inserted between the track and an Electro Code unit. The Electrified Electro Code unit converts each pulse from the Electro Code transmitter into a burst of an audio carrier at 156 Hz. Similarly, the Electrified Electro Code unit performs A.M. detection on the bursts of 156 Hz energy from the track, turning them back into dc pulses that are then passed to the Electro Code receiver. Electrified Electro Code units also have the capability to generate and transmit the 100 Hz and 250 Hz carriers used for cab signaling, and do so as directed by the Electro Code unit.

4-92

Overview of Railroad Signal Circuits

Microtrax Initially introduced by Union Switch and Signal (U.S.&S) as Microcode in the early 1980s, a redesigned product, Microtrax, became available in 1995. To overcome track storage effect, Microcode and Microtrax employ a bi-polar pulse train that is essentially a 20 Hz ac waveform. The message transmitted between the track circuit ends is modulated using a NRZ (non-returnto-zero) code with guard transitions at the beginning and ending of each code cycle. Microtrax provides continuous transmission of vital data equal to four simultaneous line circuits or bits. Microtrax can control five signal aspects and two vital auxiliary functions, consisting of 10 aspects and 1 auxiliary function, or 20 aspects and no auxiliary functions. The Microtrax circuit is designed to operate with track circuits up to 22,000 feet in length with 3-ohm ballast. Track circuits that are double this length have been successfully operated in arid regions of Australia. Protection against insulated joint failures is provided by examining the sequence of plus and minus pulses. The Microtrax unit presents a stable impedance across the track, which can be an important characteristic where Crossing Predictors or Motion Sensors are used. The bandwidth of the Microtrax receiver is very narrow; input impedance at 150 Hz is on the order of 10 ohms and may be boosted, where necessary, by placing additional inductance in the track leads. A limitation shared by all coded track circuits is the added time delay in recognizing the shunting or un-shunting of the track. To prevent spurious false occupancy indications, the Microtrax track repeater bit (functionally equivalent to the track relay) is sustained (held up) until two consecutive messages have been missed or rejected as invalid. On this basis, the minimum shunt recognition time is slightly over six seconds, and under worst-case circumstances, may be about 11 seconds, depending on when the first axle enters the track circuit with respect to the timing of the code cycle. To pick-up the track circuit (declare it unoccupied), two consecutive 6-second message cycles must be successfully received; therefore, the pick-up delay is between 12 and 18 seconds, depending on when the last axle vacates the circuit. This makes it much slower than a simple dc track circuit.

Figure 4-71 Microtrax Modulation Pattern

4-93

Overview of Railroad Signal Circuits

AC Track Circuits In General Although they are simple and reliable, and normally have fast response times, dc track circuits have several limitations that preclude their use in all cases. The most common contraindication is the use of electric propulsion for trains, and the Impedance Bonds that must be installed in electrified tracks. Additionally, track storage effect and a susceptibility to foreign dc current are important considerations. AC track circuits were first introduced around 1904, and offered a workable solution to these problems. Today, the ac track circuit remains a sophisticated and evolving technology. AC track circuits with the capacity to quickly transmit digital data to moving trains are at the forefront of contemporary track circuit development. AC track circuits posses several unique characteristics that are readily exploited for significant technical and economic advantage. The ability to readily step-up or step-down voltages with a simple, inexpensive, reliable component (a transformer) is one obvious example. Inductors or capacitors may be used for current-limiting purposes, allowing the introduction of essentially loss-less attenuators into the circuit. AC signals of differing carrier frequencies and modulation patterns may be selectively filtered using tuned circuits. AC signals may be compared against a phase reference to determine their instantaneous polarity (phase), and can thereby detect subtle differences in the circuit impedance. AC signals may be inductively coupled to or from the rails with iron or air-core transformers. Rail impedance, which is significant at frequencies even as low as 20-25 Hz, can be used to limit the effective range of audio-frequency track circuits without the necessity of insulated joints. Although the impedance of the track is higher for all ac frequencies than it is for dc energy, this has not prevented the development of a wide variety of ac signaling systems. Fortunately, increased transmission losses in the track do not necessarily require correspondingly higher amounts of power. With correctly proportioned transmitter levels and receiver sensitivities, track circuits can be successfully operated at all frequencies from dc up to 25 KHz or higher. Using a train detection carrier frequency of approximately 20 Hz and cab signaling carrier of 40 Hz, track circuits of 45,000 feet in length are now in regular service on lines in remote, arid regions of western Australia. In terms of their susceptibility to ac interference, ac track circuits are more susceptible than the most basic dc track circuits, but vary widely in their susceptibility as a function of their specific design. Classic AC Track Circuits The nomenclature used for railroad track circuits can be confusing, as it reflects a series of evolutionary changes made over a period of many decades. Perhaps the most unfortunate result of this history is that each new development was named in the most general way for the advance that it represented. In the beginning, all track circuits were dc track circuits. The use of ac energy to operate track circuits was once a novelty, thus, we are left with generic terms like: AC Track Circuits, even though this term is now normally used only to refer specifically to track circuits using power-line frequencies. 4-94

Overview of Railroad Signal Circuits

Although A.F.O.s, A.F.T.C.s, Cab Signals, Crossing Predictors, Motion Sensors, and several other types of track circuits (see below) all use alternating-current (ac) electrical signals to detect trains, these were all developed well after the original AC Track Circuit. In this section, we will be discussing the original forms of the ac track circuit, which simply detect the presence of a train on a section of track in a vital fashion, using alternating current that is usually derived quite directly from an electric utility. Since the types of ac track circuits discussed here are usually operating at exactly the same frequency carried by nearby power lines, the potential for EMC problems is significant. Attempts to use dc track circuits on electrified railways using dc propulsion systems were unsuccessful from the outset. Having a basic understanding of the simple dc track circuit, developers then proceeded to explore the possibility of using alternating current for train detection purposes in dc electrified territory. Fortunately, the construction of urban and interurban electrified railways after the year 1900 coincided neatly with the development of an ac track relay in 1904. This type of relay has a moving aluminum or copper vane similar to a disc-type induction motor (not unlike the ones found in electric utility watt-hour meters). As with the dc track relay, an upward movement of the vane closed the front contacts, indicating that the track circuit was unoccupied. When current to the coil was cut off, gravity acting on the armature or vane would return the mechanism to its de-energized position, thus opening the front contacts. The same non-welding contacts, frictionless trunnion bearings, rigid mechanical structure, generous creepage distances, and other design features commonly found in dc track relays incorporated to ensure stable and reliable operation. Using ac energy for the track circuit, and providing a receiver (an ac vane-type track relay) that is unresponsive to any level or polarity of dc, proved to be nearly as important as the concepts in Robinsons original track circuit, from both a technical and a commercial standpoint. It may also be observed that the development of ac track circuits was an early step forward in the field of EMC, having successfully solved a seemingly impossible compatibility problem between the dc propulsion current, and the signaling system.

4-95

Overview of Railroad Signal Circuits

Figure 4-72 Vane Relay (Safetran Systems)

The earliest ac track relay featured a single winding on a C-shaped stator that induced eddy currents into a flat aluminum vane attached to a rotating shaft. Shading coils were provided to create an upward torque, lifting the vane against the force of gravity when properly energized. As with the dc track circuit, a train shunting the track tended to starve the track relay of the current needed to lift the vane and keep the front contacts closed. Like the simple battery and resistor of the dc track circuit, the ac track circuits transmitter was also simple. The transmitter consisted of a step-down transformer, with deliberately poor voltage regulation characteristics, supplied by the ac mains. As with the dc track circuit, a resistor was inserted in series with the secondary of the transformer, to prevent excessive current draw when the track was shunted. The Single-Element AC Vane Relay proved quite successful, but was limited in application by several factors. First, the rail impedance of the track at 60 Hz was considerably higher than its dc resistance. Secondly, the vane relay was not particularly efficient, and consumed much more energy than its dc counterpart. This limited the maximum possible length of the track circuits. Most important, however, was the vexing problem of inter-marrying the ac track circuit and the dc propulsion return network. Facing the inevitable reality that the rails would be used as the primary return path for propulsion current, some compromise would be needed in the configuration of the track circuit. (See: Double-rail AC Vane-relay Track Circuits and Impedance Bonds, below.) Double Element Vane Relays The inefficiency of the single element vane relay could not be adequately improved upon through better materials or design techniques. The solution was to provide a Double Element Vane Relay having a second winding and core, which was physically offset from the track 4-96

Overview of Railroad Signal Circuits

winding so that the fluxes induced into the vane were in quadrature. Eliminating shading rings on stator polefaces, the new Local Winding furnished 75-90% of the energy needed to lift the vane; the Track Winding would contribute the remainder and provide the necessary offset in fluxes that generate upward torque. The double element relay was, in essence, a mechanical amplifier, phase-selective filter, and level detector, all in one package. The double element relay proved to be as immediately successful and useful as its single element counterpart. Moreover, the improved efficiency of the double element relay now placed it in direct competition with the best dc track circuits available at the time. While the attorneys battled over various track circuit installations in electrified territory, the ac track circuit was now even being applied to tracks for steam-powered trains, where no impedance bonds were required. Double element vane relays also offer an excellent release ratio, typically 75% or higher. Unsurprisingly, nearly all ac vane track relays manufactured in the U.S.A. since about 1960 are of the 2-position, plug-in, double-element type. The additional receiver sensitivity that became possible with the adoption of the double element vane relay and double-rail ac track circuit was a significant improvement in railroad signaling technology. This was important, because the performance and maximum permissible length of a track circuit are degraded by increased rail impedance, as well as by reduced ballast impedance. AC track circuits suffer at least slightly from both parts of these physical laws. Phase Shifting Further improvements in ac track circuit performance were made by adding a resonating winding to the vane relay to compensate for the increasing phase shift of long track circuits. By offsetting or correcting the phase of the local current (the current in the second winding) in proportion to the phase shift of the current from the track, workable ac track circuits of 5,000 feet in length (or more) became possible. In addition to the use of ac track circuits for basic train detection purposes, more sophisticated versions of ac relays were also developed. Similar to the polarized dc track relay, a 3-position double element vane relay was developed. By inverting the phase of the energy applied to the feed end of the circuit, the track could be used to communicate signal control information as well as detect trains. Centrifugal AC Relays It should be noted that although vane relays successfully met the demand for a dc-immune track relay, the vane relay is unsuitable for use in 25 Hz ac propulsion territory because it is not frequency-selective. Under normal operating conditions, the current flowing through the track winding of a vane relay must be of the proper level, and have the proper phase relationship in order to produce uplifting torque on the vane. Improper phase will cause the relay to drop, even if the level of current received is well above the nominal operating value. Moreover, the track and local windings must

4-97

Overview of Railroad Signal Circuits

be fed from the same source of ac energy, since any drift in frequency would cause the relay to periodically pick and drop as the two signals drifted in and out of phase with one another. 25 Hz propulsion energy may invade both the track and local windings. This may happen if induced or conducted 25 Hz energy enters the ac signal mains. Under conditions involving a broken rail or other propulsion fault, 25 Hz energy may be back-fed from the track, through the track transformer and into the track relays local winding. Acting in combination with the same 25 Hz energy appearing on the track relays track winding, the track relay will falsely pick up while the track is occupied. Countermeasures such as filtering, shielding and spatial separation of power mains have proven disadvantageous. Early in the development of vane relay track circuits, resonant L-C techniques were tried and quickly dismissed, primarily because capacitor tolerances were poor, and components varied widely with temperature and deteriorated with age. The solution was a frequency-selective centrifugal track relay that cannot be falsely picked-up by any level of propulsion current. In the first installation of this kind (circa 1911), use of 60 Hz for the track circuit frequency was selected for convenience, and without regard to the utility line interference problems that would develop years later. The centrifugal track relay uses a small induction motor driving an Archimedean governor on a vertical shaft. When the motor revolves at high speed, driven by 60 or 100 Hz track circuit current, the flyball mechanism was forced outward far enough to actuate the front contacts of the relay. And when the track was shunted, the motor produced little or no torque, slowing or stopping its rotation. If the motors track and local windings were overwhelmed by 25 Hz propulsion current, the motor might continue to turn, but only at a much lower speed. The relay was designed and adjusted so that the front contacts could not close or remain closed at less than the nominal operating speed. Centrifugal relays are also provided with an anti-reverse rotation ratchet, preventing the motor from beginning to turn opposite to its normal direction of rotation. This feature ensures that shorted insulated joints, which can supply energy of the wrong phase, cannot falsely energize the track relay. Despite its delicate construction and continuously rotating motor, the centrifugal frequency relay proved serviceable and remained the standard form of track relay in ac electrified territory for over 20 years, until supplanted by development of the coded universal track circuit in the early 1930s. Production of new centrifugal track relays was nil by 1960. The last remaining centrifugal relays, many installed in the 1930s, are just now being retired in former Pennsylvania Railroad electrified territory. Single-Rail Vane-Relay AC Track Circuits An early solution to the propulsion current return vs. signal system reliability dilemma used ac vane relays on a single-rail track circuit. Rather than isolating both rails at each end with insulated joints, only one rail would be segmented for track circuit purposes. The other rail would remain unbroken, and would be heavily bonded to return the propulsion current back to the nearest substation.

4-98

Overview of Railroad Signal Circuits

Figure 4-73 Single Rail Track Circuit

The single-rail track circuit proved to be workable, and was satisfactory where the track circuit was no longer than 500 to 1,000 feet. Longer track circuits were a problem, however. Since the single-rail track circuit has only one-half the ballast resistance of its double-rail counterpart, the track is much poorer as a transmission line. Secondly, and more importantly, the voltage drop experienced by the propulsion current in the single return rail was excessive on long track circuits. This sometimes resulted in destructive levels of propulsion current flowing through the track relay and track transformer. Single-rail ac track circuits also suffered from a safety-related shortcoming, i.e. broken rails on the return-rail side of the circuit are undetectable, due to run-around paths, which are particularly possible in multiple track territory. The single-rail ac track circuit consists of the same three essential elements: a transmitter (track transformer), a receiver (track relay), and the track. Single-rail vane-relay track circuits up to 1000 feet are preferred in dc propulsion territory on low-speed trackage where broken rail detection is not required. In comparison to other types of track circuits suitable for use in dc electrified territory, the single-rail vane-relay track circuit is the least complex, the most reliable, and offers better shunting performance.

4-99

Overview of Railroad Signal Circuits

Double-Rail AC Vane-Relay Track Circuits and Impedance Bonds Double-rail track circuits are arranged to use both rails for return of the propulsion current. This is done by placing a large center-tapped inductor known as an impedance bond at each end of every track circuit. This special inductor has windings sturdy enough to handle the large (hundreds or thousands of Amperes) dc propulsion return currents carried by the rails. The railto-rail inductance of the impedance bond is also high enough to make it look like an open circuit at ac frequencies used in railroad signaling. Pairs of these impedance bonds are used at each set of insulated joints. The center taps of the two impedance bonds are connected together with heavy-gauge cable, to provide a continuous path for propulsion return currents along the rails. Eventually, the tracks will reach a point near the propulsion substation, where the center tap(s) of the impedance bond(s) will be connected directly to the substation ground.

4-100

Overview of Railroad Signal Circuits

Figure 4-74 Single-Rail and Double-Rail Track Circuits

4-101

Overview of Railroad Signal Circuits

Figure 4-75 Double-Rail AC Vane-Relay Track Circuits

Figure 4-76 Diagram of Impedance Bonds and Insulated Joints

4-102

Overview of Railroad Signal Circuits

The purpose of the impedance bonds is to make sure that the returning dc propulsion current from the train flows freely through the unoccupied track circuits without diminishing the relatively high rail-to-rail impedance needed to prevent undesired loading or shunting of the AC track circuit. This requires that the propulsion return current be carried equally by both rails, without shorting them directly together. By using both rails to carry the propulsion return current, the double rail track circuit also has a distinct advantage over the single-rail track circuit, in that it minimizes the resistance of the propulsion return path. At the impedance bond, returning propulsion current flows from both rail connections into opposite ends of the impedance bonds center-tapped inductor. These currents are summed together, and flow out of the impedance bonds center tap. The current then passes into the second impedance bond through its center tap, and is distributed more-or-less equally into the two rails of the next track circuit. Because the currents in each rail are roughly equal, the opposing magnetic fluxes produced by two halves of the impedance bonds winding tend to cancel each other, yielding nearly zero net magnetization of the inductors laminated iron core. With little net magnetization of the core, there is also very little potential for magnetic saturation of the core material. As a result, the railto-rail inductance (and impedance) presented by the impedance bond to the tracks changes very little in response to the passage of a large propulsion return current. The most important result of this characteristic is that the rail-to-rail impedance, at the ac track circuit frequency (usually 60 Hz), remains essentially constant. If this was not the case, and the inductor began to saturate, this could result in unoccupied track circuits falsely appearing occupied whenever the train was accelerating. The net rail-to-rail impedance at 60 Hz for simple un-resonated (non-tuned) bonds is typically in the range of 0.25 to 1.5 ohms, but newer designs have increased this impedance substantially. Tuned bonds contain capacitors connected to additional windings on the core, raising the rail-torail impedance at a particular frequency or frequencies. The marriage of the vane relay and double rail track circuit was immediately successful. However, because their inventors worked for opposing commercial interests, years of patent litigation ensued, slowing development and deployment of an otherwise very successful technology. Like all other track circuits where impedance bonds are used, saturation of an impedance bond caused by interfering currents can affect the operation of the double rail ac vane relay track circuit by creating a partial shunt across the track (between the rails). All-AC Signaling Systems With the advantages presented by ac vane relays over the previous dc relays, all-ac signaling technology quickly took hold around 1905-1910. AC track and line relays had proven superior to their contemporary dc counterparts in various ways. Adding to this was the simplicity and convenience of transmitting and distributing ac energy. All-ac signal systems quickly became popular wherever ac power mains could be economically installed on existing pole lines.

4-103

Overview of Railroad Signal Circuits

The all-ac signaling era lasted into the late 1920s, when the advent of low-cost dc power supplies using copper-oxide rectifiers permitted the substitution of dc relays in line circuits and logic functions. Other factors which ended the heyday of all-ac signaling systems included the arrival of improved dc shelf-mounted relays, which were introduced in the late 1920s, and plug-in style dc relays, which arrived in the mid-1930s. Primarily due to the large size and high cost of ac line relays, few, if any, all-ac signaling systems were installed after World War II. All-ac systems also suffered from limitations related to the parasitic capacitance in cables and the inconvenience of high operating voltages (55 or 110 volts ac versus 10-18 volts dc for dc signal systems). Increasingly complex logic functions, especially stick18 functions such as route locking, demanded some means of uninterruptible power. This was easily accomplished with batteries in a dc system, but impossible (at the time) in an all-ac system.

Figure 4-77 Large Shelf Relays Characteristic of an All-AC Signaling System

Style C Track Circuits During studies of shunting performance in the 1930s, a hybrid ac-dc track circuit was developed. The simple, versatile Style C track circuit, so named simply because it was third on a list of candidates for development, uses a half-wave rectifier (a single diode) located at the far end of

18

The railroad signal term stick is generally synonymous with the digital logic term latch.

4-104

Overview of Railroad Signal Circuits

the track circuit. AC energy is applied to the track circuit, and the resulting dc current (unipolar half-way ac, really) in the rails is sensed by an ac-immune dc track relay. When the track is shunted, the diode is effectively shorted out, and the current in the rails becomes entirely ac, which cannot energize the dc track relay. This causes the relay to become de-energized and indicate that the track is occupied. Style C track circuits have been available from several manufacturers for several decades, but have limited range. The Style C circuit was once widely used as a part of a low-cost highway crossing warning system, where train speeds were 40 mph or less. The Style C track circuit is centralized, in that it only requires power at one location, and requires only one pair of track wires19. Style C track circuits are generally incompatible with Audio Frequency Overlay (AFO) track circuits and some Motion Sensors, but can be used reliably with some limited Crossing Predictor installations. Style C track circuits are generally viewed as being a source of ac interference on railroad tracks, rather than a receptor. Audio-Frequency Track Circuits So-called audio-frequency track circuits (A.F.T.C.s) are a subset of ac track circuits in general, but the term A.F.T.C. is usually reserved for modern electronic track circuits using ac frequencies within the audio band, as opposed to ac track circuits specifically using ac power from a utility or local ac power system as their operating frequency. Track circuits using audio frequencies are used in Crossing Predictors, Motion Sensors, track circuits of several different designs, and in cab-signaling track circuits. These audio-frequency track circuits use a variety of methods to couple the audio-frequency signal into the rails, including special impedance bonds, air-core loops, and direct application of the audio-frequency signal to the rails. Like most other track circuits, the A.F.T.C. relies on the same basic transmitter-shunt-receiver architecture to detect trains. The development of short-range audio-frequency overlay track circuits (A.F.O.), operating at 300 Hz to 25 KHz began in the mid-1950s, concurrent with the commercial availability of transistors. The resulting so-called jointless audio-frequency overlay track circuit is widely used for train detection (occupancy only, i.e. no broken rail detection) at highway crossings and electric switch locks. Other common applications include using these track circuits to activate 20 some hazard detectors and A.E.I. tag interrogators .

19

The diode is connected between the two rails at the opposite end of the circuit, and is installed in a rugged, weatherproof enclosure buried in the track ballast or left laying on top of it. A.E.I. stands for Automatic Equipment Identification. A.E.I. tag interrogators are systems which track the movement of railroad cars throughout the rail network. A small transponder tag is mounted on each car, and the wayside-mounted interrogator reads the information contained in the tag as the train passes by. The interrogator is activated by the presence of a train or cars on the tracks in the immediate vicinity of the tag reader system.

20

4-105

Overview of Railroad Signal Circuits

Audio-frequency track circuits are also widely used for the usual track circuit occupancy duties on rapid transit and high-speed rail systems, where the block lengths are relatively short. These applications include track circuits designed to determine the occupancy of a section of track, as well as perform other functions such as broken rail detection, and communication with the train (cab signaling). A.F.T.C.s differ from the similar Audio-Frequency Overlay (AFO, discussed in greater detail below) in that A.F.T.C. is considered a base or primary track circuit, and is generally not overlaid upon some other form of track circuit. Also unlike AFO circuits, the A.F.T.C. is designed to be capable of detecting broken rails. Base A.F.T.C.s up to 2,500 feet in length are used extensively on rapid transit lines having sophisticated automatic train control systems. Because short track circuits are necessary in order to provide a shorter headway between trains, the average length of the A.F.T.C.s used is usually on the order of 500 to 700 feet. Because the impedance of railroad track increases at all frequencies above zero Hz at a rate of roughly 3 Ohms per Kilohertz per 1,000 feet of track, the impedance of the track itself can be used as a means of limiting the effective length of an A.F.T.C. Trains shunting the track anywhere between the location of the A.F.T.C. transmitter, and the A.F.T.C. receiver, will be detected, as would be expected. But trains shunting the track a few feet (or up to a few hundred feet in some cases) outside of this area can also be detected. The detection range outside of the area defined by the A.F.T.C. transmitter and receiver is a function of the sensitivity of the track circuit (which is adjustable), the ballast resistance, and the audio-frequency attenuation characteristic of the track. Because of this attenuation characteristic, insulated joints are generally unnecessary, except where a more precise determination of a trains location is required. However, the size of the shunt detection area beyond the track wires of the track circuit is usually a minor concern. What is usually more important is the consistency of the size of this area. At higher frequencies in the range of 9 to 16 KHz, A.F.T.C.s can detect shunts that are 3 to 5 feet outside of the circuit. This is a very short length of track, and there is a correspondingly small amount of variation in this external detection distance under the influences of changing ballast resistance and other factors. At lower frequencies, both the external detection distance, and its variability, will be larger.

4-106

Overview of Railroad Signal Circuits

Figure 4-78 Impedance Bonds

Different A.F.T.C. operating frequencies are used to avoid interference between adjoining track circuits. Tuned impedance bonds can be tuned to maintain a low rail-to-rail impedance at all frequencies except those signaling frequencies for which the bond is tuned. Many jointless track circuits also employ some form of resonant shunt at each track circuit boundary to further define the boundaries of the circuit. Often this is built into the coupling network which connects the A.F.T.C. electronics to the track. The use of impedance bonds to launch and capture A.F.T.C. signals imposes significant limitations on several aspects of track circuit design. The size and thermal rating of impedance bonds can affect the operation of A.F.T.C. circuits through sometimes-subtle processes. For example, the cores of impedance bonds can be saturated by an unbalanced flow of dc propulsion current through opposite halves of a bond. This can cause the impedance bond to begin to effectively shunt the track via the reduced inductance of the center-tapped inductor in the bond. This provides a mechanism for strong interference currents (ac or dc) to saturate an impedance bond, and thereby affect the operation of track circuits. One solution to this is to use air-core inductive loops (wires) laid near the rails to couple the signaling frequency into the track. Air-core loops can eliminate the problems associated with impedance bonds such as imbalance saturation and the expense of tuned impedance bonds. Another important benefit of the air-core loop is complete galvanic isolation of the track circuit apparatus from the running rails. Unfortunately, the air-core loop is relatively inefficient as a coupling mechanism, and is thus limited in application to short track circuits (up to 1000 feet). Owing to the need for long loop length, and higher levels power, the use of air-core loops to couple audio-frequency signals to or from the rails is generally impractical at frequencies below 2 KHz. Audio-Frequency Overlay The Audio-Frequency Overlay (A.F.O.) track circuit operates in the same manner as its dc and power-frequency counterparts. The term overlay implies that the A.F.O. circuit is overlaid 4-107

Overview of Railroad Signal Circuits

upon some other form of track circuit, although it can also be used in a stand-alone fashion in some applications. This is also true of other types of circuits not normally defined as A.F.O. (e.g. Crossing Predictors and Motion Sensors). In actual practice, a half dozen or more overlaid track circuits of various types may be present in the vicinity of some complex highway crossings. A.F.O. transmitters and receivers are generally located in close proximity (less than 100 feet) to their connections to the rails, but this can be extended for some circuits with the use of special coupling circuits. These coupling circuits enable the electronics to be located many hundreds or even thousands of feet away from where their track wires are attached to the rails, thus allowing the transmitter and receiver to be placed in the same equipment shelter, even though the track circuit may be thousands of feet long. Early A.F.O. track circuits of the mid-1950s employed unmodulated carriers. The security of A.F.O. track circuits was later improved by the incorporation of Amplitude Shift Keying (A.S.K.), Frequency-Shift Keying (F.S.K.), and Phase-Shift Keying (P.S.K.) modulation techniques. More recently, digital encoding techniques have been developed to eliminate frequency coordination difficulties and reduce logistical expenses. Intelligent A.F.O. track circuits employing sophisticated digital signal processing (D.S.P.) techniques have 21 also been developed to improve the reliability of short island circuits. The frequencies used for A.F.O. circuits can range from the low tens of Hertz, to perhaps 50 KHz or more. Generally, the frequencies used are selected from a standardized set of frequencies established by the manufacturer, although some newer A.F.O. track circuits using spreadspectrum modulation techniques such as frequency hopping are difficult to describe as having a specific operating frequency. One of the most common applications of A.F.O.s is for train detection at highway crossings (grade crossings) on both electrified and diesel-electric railroads. Although they do not provide the operational benefits of Crossing Predictors and Motion Sensors, such as minimizing highway traffic delays, A.F.O. track circuits of up to 4,000 feet are common in diesel-electric territory. However, the track circuit length (approach circuit) needed to provide adequate warning time for grade crossings averages 2,500 feet or less where train speeds do not exceed 50 miles-per-hour. The vital control logic used at such grade crossings usually incorporates timers and other additional watchdogs to prevent unwanted activation of the grade crossing warning system due to failed A.F.O. track circuits. Most modern A.F.O. track circuits are often used specifically because of their generally high immunity to interference from propulsion currents and other forms of interference. In this regard they are much less susceptible to ac interference than Motion Sensors or even Crossing Predictors. However, as most A.F.O. track circuits only provide simple occupancy information, and do not measure train speed or direction, they often cannot be used as a substitute for Motion Sensors or Crossing Predictors in high-interference environments. This is because the operational behavior and performance of A.F.O.-based crossings is fundamentally different from that provided by Motion Sensors and Crossing Predictors. Furthermore, local, statewide, or national
21

An Island is a short circuit designed to detect the presence of a train within the immediate area of a grade crossing, extending for perhaps 50 feet on either side of the road itself.

4-108

Overview of Railroad Signal Circuits

policies often require new grade crossings and upgraded grade crossings to be equipped only with constant warning time equipment (i.e. Crossing Predictors). Impulse Track Circuits One substantially different type of track circuit modulation technique has been widely used outside the United States for many years. The impulse track circuit derives its name from the impulse generated when a 50 F capacitor charged to about 150 volts is repeatedly discharged across the rails. The resultant waveform at the receiving end of the circuit has a distinct asymmetrical bipolar waveshape that can be reliably detected using simple magnetic elements. The peak current flows for 1-2 ms during the first half-cycle. Because the track is a lossy transmission line (i.e. it is significantly damped), oscillation is virtually nil after only 3-4 cycles. Due to its higher rail-to-rail voltage, the primary advantage of the impulse track circuit is its excellent shunting performance in the presence of rust, sand, or other wheel-rail insulators (i.e. contributors to poor shunting). Impulse track circuits typically operate at a 4 Hz repetition rate. However, by selectively controlling the time interval between pulses, as many as 10 different codes may be transmitted using only 2 Hz of bandwidth between 3 Hz and 5 Hz. This system also offers excellent immunity to EMI because the peak rail current is on the order of 150 Amperes, generally far above the level of interfering signals produced by noise from propulsion equipment. Decoding is accomplished using pulse detectors and timers, rather than L-C filters. Lower code rates are assigned to more permissive speed commands. One disadvantage of the impulse track circuit is the need for careful insulated joint maintenance, due to the high peek voltages of this track circuit. Cab Signals Most cab signal systems use the track itself as the medium for transmitting cab signal information from the wayside to the train. Information is conveyed by a power frequency or audio frequency carrier signal applied to the track, which is periodically interrupted (100% Amplitude Modulated) at a rate of 75, 120, or 180 pulses-per-minute in a square-wave fashion (50% duty cycle). Each cab signal aspect to be displayed is assigned to a unique modulation (interruption) rate, carrier frequency, or combination of the two. This electrical signal is detected by inductive pickup coils mounted just above the rails on the front of the lead locomotive or car of the train. This coded signal is always applied to the tracks at the exiting end of the track circuit, so that the head end of the train is always moving toward the source of transmission. In order to be able to receive a cab signal indication when running in either direction on a single track, cab signal transmitters must be located at both ends of the track circuit. These transmitters are switched on selectively, depending on the direction of train traffic. Obviously, the presence of a broken rail, a false shunt across the tracks, or the presence of another train ahead in the same track circuit would tend to cut off reception of this 4-109

Overview of Railroad Signal Circuits

transmission by the approaching train. This provides the same measure of safety as conventional track circuits and wayside signals, in that a dark signal aspect (no indication) must be treated in the same way as the most restrictive signal indication (STOP).

Figure 4-79 Cab Signaling and Automatic Train Control were Well-Established Technologies by 1940

In general, cab signal aspects are controlled by the next wayside signal ahead of the train. Railroad signal personnel refer to this as the next signal in advance of the train. Because systems using this simple modulation scheme are generally limited to a small number of code rates and associated aspects (2-6), it becomes impractical to fully replicate in the cab the 9-17 different aspects that can be given by a multiple-headed set of wayside signals. After being inductively picked up by the receiver coils located ahead of the first axle on the locomotive, the incoming signal is then amplified and decoded to control the in-cab aspect display. An upward or downward transition of the cab signal aspect may be used to control an audible warning (a small whistle in the cab). A downward transition in cab signal aspect may also trigger a penalty brake application if not suppressed by a manual brake application or acknowledgement from the engineman within 5 to 8 seconds. This penalty braking application consists of a full service brake application that cannot be released until the train has been brought to a complete stop. In more sophisticated Automatic Train Control (A.T.C.) systems, the speed of the vehicle is compared to the allowable speed for a given aspect. If the vehicles speed exceeds the allowable limit, the engineman must reduce the speed or manually make a full service brake application in 4-110

Overview of Railroad Signal Circuits

order to avoid the penalty braking. Some systems even employ a declining speed profile, whereby the penalty braking application is forestalled so long as the trains speed gradually decreases in accordance with a predetermined braking curve. Early coded cab signal receivers of the 1927-1950 era employed two-stage tube amplifiers to amplify the signal from the inductive pick-up coils. Frequency-selection, demodulation, and decoding of the carrier were accomplished using tuned L-C circuits driving the coils of dc relays through copper-oxide rectifiers. Transistor amplifiers were introduced as early as 1953 in carborne applications, and microprocessor-based circuitry has since eliminated most of the bulky, heavy components. The on/off duty cycle of the modulation is nominally 50%, with code (pulse) rates in the range of 0.83 to 21 Hz. However, the standard code rates of 75, 120, and 180 pulses-per-minute (1.25, 2.00, and 3.00 Hz, respectively) are the most common. Carrier frequencies can range from 40 Hz to as high as 16 KHz. As a general rule, higher code rates are assigned to higher speed limits, and the decoding circuitry is arranged so that the lowest recognized code rate governs train speed, in order to prevent the cab signal system from displaying a more permissive aspect than intended, due to harmonics or the presence of interfering electrical signals on the rails. 40 Hz carrier has been successfully used on track circuits of 12,000 feet or more. The lower reactance of railroad track at low frequencies is the primary motivation for adopting 40 Hz as a cab signal carrier frequency. Under exceptionally dry track conditions, as found in Western Australia, 40 Hz cab signal is successfully operated over track circuits up to 45,000 feet long. Unfortunately, 40 Hz track circuits are plagued with interference from many sources, especially localized residual magnetism in the rails, joint bars, and tie plates, which generate low-frequency electrical noise when the cab signal sensing coils on the front of the train pass over them. For the same reason, attempts in the 1950s to detect and decode pulses on rate-coded DC track circuits proved unsuccessful. 60 Hz carrier, with a minimum carrier level of 0.5 Amps at the entering end of the circuit has been successfully used on track circuits up to 11,000 feet in length. Unfortunately, 60 Hz cab signal systems have proven increasingly difficult to maintain due to ever-worsening interference from utility transmission and distribution systems. One two-aspect 60 Hz train control system was successfully converted to use a 100 Hz carrier frequency in the mid1970s. However, the increased reactance of the track at 100 Hz effectively reduces the maximum operable length of track circuits, making the transition from 60 to 100 Hz difficult and expensive. 100 Hz carrier, with a minimum carrier level of 2.0 or 3.0 amps at the entering end of the circuit, is the most widely used system today. Track circuits up to about 7,500 feet in length are common. To avoid heterodyning with the fourth harmonic of 25 Hz propulsion energy, a carrier frequency of 91-2/3 Hz is used in areas where dual AC/DC propulsion systems exist. Code rates of 75, 120, and 180 pulses-per-minute are most commonly used. Carrier frequencies of 75, 83.3, or 90 Hz are countries used in foreign countries having 50 Hz utility systems. Carrier frequencies lower than 60 Hz are generally impractical in electrified territory due to the low rail-to-rail impedance of un-resonated impedance bonds. Frequencies higher than 100 Hz usually dictate use of low current levels. 125 Hz cab signal carrier has been used in Europe. Certain ATC systems in Asia have also employed single-sideband 4-111

Overview of Railroad Signal Circuits

transmission techniques, which are synchronized to the 60-Hz propulsion current, that operate in the range of 720-1020 Hz. 250 Hz carrier is now becoming more common for high-speed rail applications. One version provides a supplemental 250 Hz carrier interspersed with the 100 Hz to provide additional combinations. This system can display nine different signal aspects.

Obviously, using a power-line-related frequency (the power-line frequency or one of its harmonics) as the operating frequency for a track circuit is likely to increase the chances of electromagnetic incompatibility with any unintended sources of ac energy. Surprisingly, 60 Hz track circuits of various types have still been developed, and continue to be used in some limited parts of North America. One of the most glaring examples of this type is the use of 60 Hz Cab Signaling. Although there are some features of this approach to cab signaling that may appear economically attractive, the downsides, including its susceptibility to 60 Hz ac interference, greatly outweigh any apparent advantages. Audio-Frequency Cab Signaling Although the carrier frequencies of 40, 60, 100, and 250 Hz used in cab signaling are clearly within the audible range of human hearing, they are for some inexplicable reason considered separately from the audio-frequency carriers used in specialize types of railroad cab signaling. The most common audio-frequency carriers for these specialized cab signaling systems in the U.S. are typically 990, 2340, 4550, and 5525 Hz. These are typically used on light-rail electricpropulsion rapid-transit systems. The minimum rail current sensitivity for the carborne receivers used in such systems is typically 150 to 300 milliamperes. Code rates for cab signal systems using audio-frequency carriers typically range from 75 to 1260 pulses per minute (1.25 to 21 Hz). Due to limitations imposed by the resonant circuits in the impedance bonds, there is a practical limit of about six or eight code rates (aspects) per carrier in this type of system. On one rapid transit system, 12 aspects are provided using 4550 and 5525 Hz carriers. Worldwide, audio-range carrier frequencies of 1700-2600 Hz, 4-6 KHz and 9-16 KHz are commonly employed. Frequency-shift keying (F.S.K.) has been used on recent audio-frequency cab signaling systems operating in the frequency range of 9 to 16 KHz. Given the generally favorable track circuit parameters (high ballast resistance and very short track circuits) found on modern rapid transit systems, an 80-bit digital message can be transmitted to the train at a rate of three times per second over track circuits of up to 1000 feet in length. Phase-Shift Modulation and Pulse-Coded Modulation techniques have been used to a limited extent in the past. Various other frequency modulation techniques have also been studied, but have not been implemented as yet. Because of the tracks poor transmission characteristics (primarily its limited bandwidth), sophisticated techniques such as Trellis Coded Modulation, now common in voice-band telecommunications, have not been applied to this form of signaling to date. In general, audio-frequency cab signaling in relatively immune to ac interference, due to its use of relatively high carrier frequencies. The short track circuit length used also helps minimize the potential for inductive coupling. 4-112

Overview of Railroad Signal Circuits

Grade Crossings In any country with a well-developed rail network, situations will inevitably arise where the railroad needs to cross paths with another mode of transportation. In many cases, this involves a grade separation, wherein bridges, tunnels, and underpasses or overpasses are used. The primary advantage of such grade separation is that it permits an uninterrupted flow of traffic on both transportation systems. However, the construction and maintenance of these structures is costly. Bridges routinely cost more than $1 Million (U.S.), and still require periodic inspection, painting, and other maintenance. The obvious and much cheaper alternative to such divided-elevation solutions is the so-called Grade Crossing. Known variously as Highway-Rail Intersections, At-Grade Crossings, Road Crossings, or Highway Crossings, the most common term for them is simply grade crossing. This includes crossings designed for pedestrian traffic, as well as the more common version designed primarily for automotive traffic across the railroad. At such grade crossings, electric and electronic warning systems are installed which permit an uninterrupted flow of train traffic - by interrupting the highway traffic only when necessary as the trains are passing. Approaching trains are given priority to use the crossing at any time, in exchange for unrestricted automotive traffic flow at all other times. Although this might seem to unduly favor the passage of train traffic, the relative braking distances of trains and automobiles dictate that it must be so. Based on the principle that the railroads were there first and own the land, Highway Crossing Warning System (H.C.W.S.) improvements and maintenance are generally funded by state transportation authorities. The state authorities determine where H.C.W.S. improvements are needed and what type of equipment is to be installed. State authorities may also pay the railroad an annual fee to maintain the installations. But in the United States, most of the capital funding for H.C.W.S. improvements ultimately comes from annual federal transportation grants. Grade crossings work well for low-to-medium train and highway traffic densities. However, when the density of either form of transportation reaches the point that the highway traffic interruptions become intolerable, some other solution such as grade separation is usually employed. This is often the case for heavily-traveled highways, or freeways crossing railroad lines, and also for some rail transit systems. Of the roughly 265,000 public grade crossings in the United States, about 65,000 of them are equipped with some form of crossing warning system. With the exception of a handful of crossings which are controlled by manual or experimental means, each crossing equipped with warning devices employs some form of track circuit for train detection, just as a signal system does. Even on railroad tracks with no other form of railroad signaling (so-called dark territory), it is still likely that you will find track circuits for highway crossing warning systems. Highway crossing warning systems are a large and growing part of the signaling systems on nearly every railroad in the United States. The evolution and deployment of warning signals in the form of flashers, gates, bells, and other devices began prior to World War I. Standardization of the warning device form and function was resolved in the 1930s. This long history has left us with warning systems that have several anachronistic characteristics. 4-113

Overview of Railroad Signal Circuits

For example, the height specified for the signs and lights at a grade crossing has its roots in the height of a mans eyes while seated on a horse. Fortunately, this height also works well for modern highway traffic in most cases. And the alternating flash of the lights at a grade crossing is the modern incarnation of the so-called wig-wag device, in which a steadily-lit red lamp was swung back and forth over the crossing, to simulate the motion of a red kerosene lantern in the hand of a train crewman standing at the crossing. Many such historical features remain, even though many of these devices are now controlled by modern digital technology.
Crossing Predictors

The basic purpose of a Crossing Predictor is to help avoid accidents involving trains and automotive traffic, and to minimize automotive traffic delays by providing a consistent amount of warning time prior to the arrival of a train at a grade crossing. By minimizing excessive warning time, there is less temptation for the motoring public either to try and beat the train in order to avoid a long wait, or to simply ignore the warning devices and cross the tracks while a train is approaching. The history of the Crossing Predictor starts in the late 1950s, when the Southern Pacific Company - better known as the Southern Pacific Railroad - contracted with the Stanford Research Institute in Menlo Park, California, to investigate the feasibility of using railroad track as a transducer to measure a trains position and velocity, and feed this information into a simple analog computer which would calculate the amount of time remaining before the train reached the crossing. In this way, a system could be constructed which would activate the warning devices (gates, lights, bells, etc.) at a railroad crossing at a consistent interval prior to the trains arrival at the crossing regardless of wide variations in train speed. The research proved fruitful, and the Southern Pacific Co. then sought a company to turn this basic research into a viable product. After a few more years of research and development, the first commercial Grade Crossing Predictor was produced by the Marquardt corporation in 1962. Known as the Model 300 Grade Crossing Predictor, it was built using discrete transistors and components mounted on a series of circuit cards plugged into a chassis with a hand-wired backplane. Even using solid-state construction, it still consumed more than 100 watts of power, and two such units were required at each crossing for each track. Advances in electronics technology have greatly increased the performance and reliability of modern Crossing Predictors, while significantly reducing their power consumption, but many Model 300s are still in service across the United States - well beyond their intended 20-year service life. The earliest crossing warning systems were constructed using three simple track circuits combined with some relay logic, and suffered from several shortcomings. One of the most annoying to motorists was the inability of the system to respond quickly to a train that had slowed to a stop while approaching the crossing. This could cause the gates to remain down, with no approaching train visible. These systems also had no way of adapting themselves to trains of different speeds, and warning times varied widely. Although there had been earlier attempts at constant warning time systems for grade crossings, some involving complicated trap circuits with numerous relays, timers, and track circuits, Crossing Predictors were the first technology to provide such an accurate and consistent warning 4-114

Overview of Railroad Signal Circuits

time for trains of widely-varying speeds. In this context, widely-varying refers to the variation in speed from train to train, with the speed of each train being essentially constant while it is within the approaches to a grade crossing. The underlying concepts of the Crossing Predictor are relatively simple. First, always remember that the axles of a train provide a short or shunt across a set of railroad tracks, and that the measured electrical impedance of a railroad track circuit, as measured from one rail to the other, depends on the distance between the axle shorting the rails together, and the point at which the measurement is made. Second, a Crossing Predictor measures both the magnitude of the track circuit impedance, and the positive or negative rate-of-change of that impedance with respect to time. By dividing the magnitude of the track circuit impedance by its time-rate-of-change, the point in time at which the impedance will reach zero (or a minimum) can be calculated. This will occur when the first axle of the train reaches the point at which the measurement of track circuit impedance is made. This measurement is usually made at the edge(s) of a grade crossing. By continuously calculating the amount of time remaining before the first axle of the train reaches the point of measurement (called the feedpoint), and comparing it to the amount of warning time desired, a decision can be made as to when to activate the crossings warning devices. Whenever the amount of time remaining is less than or equal to the desired warning time, the warning devices (gates, lights, bells) are activated.

Figure 4-80 GCP Block Diagram

4-115

Overview of Railroad Signal Circuits

A negative rate-of-change in track impedance (impedance diminishes over time) indicates that a train is approaching the crossing. Conversely, a positive rate of change signifies that the train is moving away from the crossing. And the magnitude of the impedance is proportional to the position of the nearest axle of the train within the approach. Thus, the exact distance between a crossing and an approaching train can be measured, as well as the trains velocity. Using the simple relationship of time, distance, and velocity: Time = Distance / Velocity, we can use these electrical track measurements to compute the amount of time required for an approaching train to travel the remaining distance to the crossing. By continuously comparing the results of this computation with the amount of warning time desired, we can easily determine when the train is warning time seconds away from the crossing. When it is, the crossings warning devices will be activated by the Crossing Predictor. Ultimately, the decision to activate the crossing consists of a single binary output. This output is normally energized to around 12 volts dc when there are no trains approaching, and is used to drive the coil of a relay or some other device in the equipment shelter at the grade crossing. Enough current drive is available to actuate a standard general-purpose railroad signal relay, which has a coil resistance of about 500 Ohms. Whenever the crossings warning devices are to be activated, the Crossing Predictor will de-energize its primary relay drive output which initiates a series of actions by the warning devices (gates, lights, bells, etc.) at the crossing. Although Crossing Predictors (and Motion Sensors) are primarily designed to detect trains approaching a grade crossing, they must also correctly handle stationary trains which are blocking the road or highway. Therefore they also contain an additional circuit (usually consisting of an audio-frequency overlay or AFO) known as the island circuit. The purpose of the island circuit is to keep a grade crossings warning devices activated any time that a train is occupying (or very close to occupying) the grade crossing. Just like any other track circuit used to determine track occupancy, this island circuit must respond to the presence of a 0.06-Ohm22 or lower resistance connected between the rails anywhere within the track circuit. The island circuit has absolute control over the grade crossing warning system, and will activate the crossings warning devices when the island is occupied by one or more axles of a train. This is sometimes known as a positive island ring. Most railroads also include roughly 50 feet of track on either side of the road as part of the island circuit, but may increase this significantly when the railroad and highway intersect at anything other than a right angle, or if there are other forms of visual obstruction which could prevent someone from actually seeing an approaching train. The island circuit is created by separating the points at which the transmitter and receiver of the Crossing Predictor attach to the track at the crossing, and using these track wires for both the Crossing Predictor and the island circuit. Since the Crossing Predictor and the island circuit use different operating frequencies, they are able to share these wires easily.

22

Many railroads, especially passenger transit lines, set their island circuits to be even more sensitive than the 0.06-Ohm standard required by the FRA. Settings of 0.12 or even 0.25 Ohms are not uncommon.

4-116

Overview of Railroad Signal Circuits

Crossing Predictors (and Motion Sensors) use their island circuits to determine when the last car of the train has cleared the crossing. When the island circuit becomes occupied, and then un-occupied after the Crossing Predictor has predicted the arrival of a train at the crossing, the warning devices at the crossing should then return to their un-activated states. Again, this seems like a simple task. But consider the following example: what happens when a Crossing Predictor has predicted the arrival of a train at a crossing, and has activated the crossings warning devices, and then the train slows to a stop before reaching the crossing? When should the gates go up? When do the gates actually go up?

Figure 4-81 Track a Island, Showing Track Wires

The actual answer to this last question is that when a Crossing Predictor has predicted, but is no longer seeing inbound motion, a timer begins to run. Should this timer complete its timing cycle before inbound motion is seen again, the crossings warning devices will be instructed by the Crossing Predictor to shut off. This interval is known as pickup delay, in reference to the delay which occurs before the railroad signal relay controlled by the Crossing Predictor picks up, i.e. becomes energized, after the train has stopped within the approach. These sorts of timing cycles are found throughout the internal operations of Crossing Predictors. In fact, you could think of a crossing predictor as a state machine controlled by an ac impedancemeasuring system, and equipped with many timers. These timers are usually used to keep the Crossing Predictor in various operating states for a period of time, generally just after something significant has occurred, in order to keep the crossings warning devices activated whenever they must be. With respect to ac interference, it is often the length of these timers rather than the persistence of the offending ac energy, which keeps the gates down at grade crossings for long periods during intermittent ac interference.
Motion Sensors

Following the advent of the Crossing Predictor, it was recognized that not all of the advantages of a constant warning time system were needed at all grade crossing locations. For example, a grade crossing which sees only through train moves (the train never stops within the approaches to the crossing) at or near the track speed limit (common in commuter rail 4-117

Overview of Railroad Signal Circuits

operations) could be served equally well by a Crossing Predictor, Motion Sensor, or simple ac or dc track circuits. In each case, the warning time would be roughly the same. But for trains which might start and stop within the approaches to a grade crossing, Motion Sensors offer a significant advantage over conventional ac or dc track circuits, in that they will allow the crossing to clear up shortly after a train stops while within either approach to the crossing, provided that no part of the train is obstructing the crossing itself (occupying the island). And if the train began to move towards the crossing again, a Motion Sensor can activate the crossings warning devices again, generally providing much more warning time, and much less train delay than would be required by conventional dc or ac track circuits. Unlike Motion Sensors and Crossing Predictors, crossings controlled by conventional dc or ac track circuits may wait many (15-20) minutes before deactivating the crossings warning devices in response to a stopped train on an approach - if in fact they ever do. And when the train started moving toward the crossing, it would have to slowly approach the crossing prepared to stop just short of the island, and creep up onto the island circuit. Once the train entered the island circuit at the crossing, the crossings warning devices would be activated, and once the gates were lowered (or the highway traffic was stopped by the lights and bells alone) the train could proceed across the crossing.

In contrast, a motion sensor would activate the crossings warning devices a few seconds after the train resumed moving towards the crossing, allowing the train to move towards the crossing without stopping. Obviously this is preferable to the operation of a crossing controlled by conventional dc or ac track circuits, and this is often one of the motivations for railroads to use Motion Sensors as a replacement for conventional dc or ac circuits at grade crossings. Electrically speaking, Motion Sensors detect the presence of trains using the track as a transducer in exactly the same way as Crossing Predictors do. What differs is the type of processing that is done on the train position and train velocity data generated. In general, a Motion Sensors decision to activate the crossings warning devices is based almost solely on the trains velocity. If the trains speed towards the crossing is at or above some minimum threshold speed, sometimes called the minimum motion threshold, then the crossings warning devices are activated. Otherwise, they remain un-activated. The only thing that complicates this process somewhat is that the minimum motion threshold can be made a function of any other system variable or input to the system, such as the trains position. This is done to prevent extremely long warning times from occurring with trains moving very slowly towards the crossing from the far end of an approach. Motion Sensors, Crossing Predictors, and AC Interference Among all of the different types of track circuits, Motion Sensors and Crossing Predictors are the two most sensitive systems with regard to ac interference, due to the fundamentally different way that these devices work, as compared to almost all other track circuits. While most circuits transmit a signal at one end of a track circuit, and receive it at the opposite end of the track circuit (or fail to receive it when the track is shunted), Motion Sensors and Crossing Predictors

4-118

Overview of Railroad Signal Circuits

are different. Their operation is controlled by the changes in the electrical impedance of the track circuit that they are connected to. A typical track circuit is like a communications link between two points on the track, which works when the track is unoccupied, and stops working when the track is shunted. In contrast, Motion Sensors and Crossing Predictors are more like an impedance meter which measures the electrical characteristics of the track, as viewed from a particular point. The measurement made by this impedance meter is then run through some additional processes to determine if there is an approaching train. Anything which interferes with the ability of a Motion Sensor or Crossing Predictor to accurately measure the impedance of the track circuit can cause the system to falsely activate the warning devices at a crossing, even though there is no train approaching. Since vitality requires that these systems be capable of detecting any internal or external conditions which could prevent the system from providing adequate warning time, they contain a series of processes for detecting internal and external failures. Motion Sensors and Crossing predictors are designed to monitor these internal and external conditions every few seconds via a built-in self-testing or selfchecking process. The presence of any apparent vital failure is sufficient cause for the crossings warning devices to be activated. The system must take a fail-safe action when any vital failure is detected. In one sense, the safest crossing is one whose arms are continuously in the lowered (horizontal) position, with the lights and bells operating. Of the two systems, Motion Sensors are more sensitive to ac interference than Crossing Predictors, due to their less-restrictive requirements for activating the crossings warning devices. Specifically, a Motion Sensor only has to detect something that looks like a sufficiently high rate of inbound motion in order to activate the warning devices. In contrast, a Crossing Predictor must see a sufficiently high rate of inbound motion with respect to the trains apparent position. This additional requirement reduces the sensitivity of the Crossing Predictor to ac interference, as compared to a Motion Sensor. Unwanted ac electrical energy which is coupled into the rails through magnetic induction or other processes is often present, but seldom a problem. Generally this type of induction affects both rails equally, raising their ac potential with respect to ground. But since Crossing Predictors and Motion Sensors are transformer-coupled to the rails, and use rail-to-rail (differential mode) 23 voltages to detect trains, this common-mode voltage is invisible to these systems . Most forms of ac interference can only occur when interfering voltages are present in the form of rail-to-rail (differential-mode) potentials. This usually happens when some of the common-mode interference is converted into differential-mode by an impedance imbalance between the two rails of the track. The kinds of problems that this can cause when it occurs will be discussed in greater detail in the Investigation and Diagnosis chapter. However, the most common symptom of ac interference is a false activation of the crossings warning devices, which is
23

One exception to this generalization would be when the rail-to-ground potential rises enough to activate the protective devices (Lightning Arrestors) which then shunt excessive rail-to-ground potentials to earth. This requires a few hundred volts of rail-to-ground potential. Another exception would be when the dielectric breakdown voltage of the wiring or components used in railroad signal equipment is exceeded. This requires a few thousand volts.

4-119

Overview of Railroad Signal Circuits

simply called a false activation by the FRA. This occurs when the presence of the unwanted ac energy interferes with the proper operation of the train detection equipment, either by simulating the electrical effects of an approaching train, or by simulating the symptoms of an internal or external vital failure. Manifestations of this sort of interference can and do vary widely, both in type and severity. In the most extreme cases, the warning devices at a grade crossing will remain activated continuously, and there will have been visible permanent physical damage inflicted upon the equipment itself. More often, ac interference is noticed when the warning devices at a grade crossing occasionally go into operation with no trains present. Not every crossing which occasionally experiences a false activation is experiencing ac interference. All too often, these problems are attributed to ac interference long before a proper investigation into the cause has been made. There are many other root causes that can produce the same end results, i.e. false activations and long warning times. Even though it may be a nuisance, it is still fortunate that the effect of ac interference on Crossing Predictors and Motion Sensors is generally to activate the warning devices at a crossing when they should not be activated (a right-side failure), rather than to prevent the warning devices from being activated when they should be (a wrong-side failure).
Conventional or Stick Grade Crossing Warning Systems

In the simplest type of grade crossing warning system, track circuits are constructed on both sides of the highway to detect the presence of a train as it approaches the crossing. Unsurprisingly, these track circuits are known as the approaches of the crossing. These approaches begin approximately 50 feet from either side of the highway pavement, and extend away from the crossing far enough to provide adequate warning time for train moving at or slightly above the district speed limit for those tracks. Filling the gap between these two approach circuits is a third very short track circuit known as the island, which spans the highway itself. By detecting the occupancy of each of these three track circuits, and using a vital logic circuit 24 known as a stick , the warning devices at a grade crossing can be activated whenever either of the two approaches has been occupied. Furthermore, the warning devices will be turned off whenever either of the following two things happens: 1. The train completes its movement across the crossing, with the island circuit first becoming occupied and then unoccupied, or 2. The train stops and backs away from the crossing before reaching it, leaving the approach circuit, which starts a timer of several minutes duration. In this case, the crossings warning devices will not be turned off until the time delay has expired. The length of the approach circuits is determined by the maximum authorized train speed, and is calculated so as to provide 20-30 seconds of warning time before the train arrives at the crossing. One or more track circuits may be located within each approach, with their outputs combined to form a single approach circuit, depending on the type of system and the layout of the tracks.
24

The railroad signal term stick is generally synonymous with the digital logic term latch.

4-120

Overview of Railroad Signal Circuits

In normal operation, the warning devices are activated as soon as the train enters either approach, and remain so until the rear end of the train has cleared the crossing. This technique is satisfactory where all trains approach the crossing at or near the speed limit without stopping. Trains approaching the crossing at less than the maximum speed will, of course, receive more than the designed warning time. Trains traveling at half of the speed limit will produce warning times that are twice as long as a speed-limit train move will. For long heavy trains that are just starting to crawl out of a railroad yard, the warning times provided by this technique can quickly run into the minutes and tens of minutes. This can begin to impact the safety of the crossing, as the credibility and effectiveness of the grade crossing warning system is compromised when warning times are perceived as excessive or erratic by the motorist. The case of a slow but accelerating train approaching a grade crossing is particularly dangerous, because the train may have reached a moderate speed by the time it reaches the crossing, and this may coincide with the limit of a motorists patience. The results are too often tragic. The impatient motorist may be tempted to ignore the warning flashers and bells, and may drive around the crossing gates (motor-operated physical barriers) at the crossing, just in time to discover that there really was a train coming. In other cases, the grade crossing warning system may be activated by trains that wind up stopping within one of the approaches before reaching the crossing. This can happen when trains stop at wayside signals. As trains may be required to wait for times ranging from seconds to hours, a grade crossing warning system that uses such simple logic must rely on its time-out timers to restore highway traffic flow.
Highway Crossing Gate Mechanisms

Highway crossing gate mechanisms are another type of signal equipment commonly found on mainline, commuter, and light-rail lines. Although an electro-hydraulic gate mechanism was marketed in the U.S. during the 1960s, and many are still in service, nearly all gate mechanisms sold today employ dc motors and mechanical gearboxes. Many early gate mechanisms installed prior to 1950 were in fact reworked semaphore signal mechanisms. All gate mechanisms use motor power to drive the gate arm upward to the vertical (90-degree or clear) position. Once in the vertical position, the gate arm is held by an electro-mechanical hold-clear device that prevents reverse (downward) rotation of the motor armature. This is necessary because the arm and counterweight assembly (if any) are balanced so that the arm will gently descend to the horizontal position by gravity if unrestrained. The gate control voltage consists of 12 volts dc at a few milliamperes (roughly 35 mA), and is what powers the hold-clear device. When this voltage is turned off by the crossing warning system, or disappears due to a complete failure of electrical power, a sequence of events takes place in the gate mechanism. During the normal operation of a gate mechanism, the gate spends most of its time in the vertical position, locked in place by the hold-clear device. The arm is vertical, the warning bell at the top of the mechanism is silent, and the lights are not flashing. Perhaps surprisingly, the attached bells

4-121

Overview of Railroad Signal Circuits

and lights are not controlled by the gate mechanism itself, but by the electronics of the warning system in the signal equipment enclosure. When the crossing warning system is activated, the flashing light units on the mast and gate arm immediately begin flashing, and the bells immediately begin ringing. A 3.0- to 3.5-second delay is then provided between the commencement of flashing and the descent of the gate arm. This is done to lessen the risk of motorists hitting or stopping beneath the gate arm. The hold clear is then de-energized, releasing the shaft of the dc permanent-magnet motor. The gate arm immediately begins descending to the horizontal (0 degree) position. To counteract the effects of unfavorable winds acting on long gate arms, the gate motor is used to drive the gate arm downward for the first third (or more) of its descent. This accelerates the gate arm more quickly than gravity alone would do, which shortens the amount of time required to reach the horizontal position. The gate arm then coasts for roughly the middle third of its descent, and is dynamically braked to a stop during the final third. This braking is done by using the gate motor as a generator, and dissipating the power generated in an adjustable resistor in the gate mechanism. The adjustment of this resistor controls the amount of braking applied, and a mechanical dashpot or buffer assembly stops the arm at the lower limit of its travel. Once the train has passed, and the gate arm is to be raised, the motor is used to drive the gate arm back to its vertical position. The lights and bells are left flashing and ringing until the gate arm is within a few degrees of vertical. Although gate mechanisms must be instantly reversible from any position of ascent or descent, not all railroads control them in this way. Some subscribe to the philosophy that a gate mechanism which reverses its direction in mid-travel is too confusing to the motoring public, and so they add logic to assure that any movement, whether towards the vertical or horizontal position, will always be completed. Wood, fiberglass, and aluminum gate arms are lightweight, and have frangible or spring-return connections to the gate mechanism. This is done to prevent the unnecessary breakage of gate arms by highway traffic. Even so, gate arm replacement is a frequent maintenance chore at some crossings. The longest gate arms typically used are approximately 40 feet in length. Articulated gate arms may be used where clearance to overhead power lines or other structures is a concern. Gate mechanism motors and control circuits are designed to withstand stalling of the arm, which may occur if the arm is obstructed or broken. Some gate mechanisms also have a second output shaft for operating a short pedestrian gate arm that blocks the sidewalk at the same time as the main arm blocks the road. The vast majority of highway crossing gate mechanisms installed in the U.S.A. today are of the half-barrier type, so-named because the gate arm blocks only one-half of the roadway. As this makes it possible for traffic to swerve around the gates, the use of traffic-obstructing center dividers and so-called four-quadrant gate layouts is becoming increasingly common. Specialized vehicle arresting barriers or nets have recently been tested in limited trials. Most of them aim to prevent a speeding vehicle from crashing through the gates and obstructing the 4-122

Overview of Railroad Signal Circuits

tracks. Unfortunately, some of these require equipment to be located along the roadway at some distance from the grade crossing, limiting their applicability to most crossings. Many level crossings (grade crossings) in the U.K. and elsewhere employ manually operated gates that cover the full length of the crossing. Although manually operated crossing barriers have a long heritage in the railroad industry, the associated labor cost is making this practice largely a thing of the past.

Figure 4-82 Typical Half-Barrier Highway Crossing Gate Mechanism

Highway Crossing Flashers

Highway crossing flashers are universally applied to all highway crossings equipped with warning devices other than a sign. Curiously, the term crossing protection is reserved for the familiar crossbuck railroad crossing sign placed at all grade crossings, with all of the electric or electronic systems referred to as crossing warning systems. At any crossing in the U.S.A that is equipped with some form of crossing warning system, you will find highway crossing flashers. A highway crossing equipped with flashers will normally have between 8 and 50 flashing red lights mounted on gate arms, masts, or cantilever structures at various points around the crossing. Typically, this would consist of two mast-mounted sets of four lights each, plus at least two flashing lights per gate arm (if gate mechanisms are used). Gate mechanisms are also required to have a tip light, which resembles the flashing lights on the gate arm, but remains continuously lit whenever the warning devices are in operation. The mast-mounted incandescent flashing light units use a 10-volt, 18-25 watt precision signal lamp, a concave mirror, and a colored roundel. LED-based light units having similar electrical ratings are now becoming widespread as a replacement for incandescent bulbs, as they offer 4-123

Overview of Railroad Signal Circuits

much longer lifetimes, graceful degradation in the event of damage, and a very intense and colorsaturated source of light. All modern flashing light units are either 8-3/8" or 12" in diameter. Lighting units are paired to flash alternately using a three-wire circuit. Voltage from a battery or ac transformer is applied to the ends of the flashing light string, and a shunt is alternately applied across each bulb (see below).

TIP LIGHT

BATTERYFLASHER RELAY

BATTERY+

Figure 4-83 Schematic of Lights and Shunts

In this way, the failure of the flasher relay can, at worst, result in both bulbs being at least dimly lit. The flashing lights at crossings are usually powered by 10-Volt ac power from a lighting transformer, with a provision for automatic switch-over to 10-Volt dc power from the back-up batteries when ac power is interrupted. Proper focus, beamspread, and maintainer access are important issues in light unit performance. FRA rule 234.221 requires that all lamps in a highway crossing warning system be energized at no less than 85% of the lamps rated voltage. This means that the battery back-up must be able to supply at least 8.5 volts to each lamp.
Crossing Bells

Highway crossing flashers are often augmented by bells to warn pedestrians. Bell mechanisms normally consist of a solenoid-operated mechanical gong driven by 10 volts dc. In contrast to much of the other equipment used at grade crossings, bells are considered non-vital. That is, their 4-124

Overview of Railroad Signal Circuits

proper functioning is not automatically checked by the system, and they are not designed to be fail-safe. Due to long-term reliability problems, electro-mechanical bells are gradually being replaced by electronic bells. Electronic bells use a weatherproof speaker connected to an amplifier and bell synthesizer. With no gross mechanical moving parts, these bells are less subject to problems caused by ice, rust, worn contacts, or lack of lubrication, but are still considered non-vital. A relatively new type of audible warning device is the stationary directional horn. Mimicking the sound of a locomotives horn, solid-state horns attached to the highway crossing signal have been investigated as an alternative to late-night locomotive horn blowing at crossings in residential areas. To-date, application has been limited to a few demonstration sites. FRA regulations require the sounding of a locomotives horn at all highway crossings, except where four-quadrant gates or other special protective features have been installed. Highway Crossing Traffic Preemption Highway crossing traffic preemption is the practice of interconnecting the grade crossing warning devices with the highway traffic signal lights. This is often done where a grade crossing is located close to (or even within) a highway intersection. The overall objective is to coordinate the operation of the traffic lights with the crossing warning system so that both systems are more effective. 1. The specific goals of traffic preemption vary with the location and design of the intersection and grade crossing, but they can include: 2. Prevent cars from becoming trapped on the grade crossing by a red traffic light 3. Prevent gates from being lowered on top of trapped cars (this helps reduce gate arm breakage and claims for automobile damage) 4. Prevent motorist confusion which can lead to accidents (e.g. it is confusing when motorists are given a green traffic light to proceed into an intersection which is blocked on the far side by an occupied grade crossing) 5. Prevent traffic tie-ups caused by intersections filled with cars that were trapped by a green traffic light and an occupied grade crossing. The form of the interface between the crossing warning system and the traffic lights can vary, but it usually takes the form of a set of dry relay contacts, which are provided to the local traffic engineer by the railroad. This relay is located in the railroad signal enclosure at the crossing, and the traffic engineer is allowed to use the contacts in any way desired. Unlike most railroad signal wiring, which carries low-voltage dc current, most traffic light control systems use 120-Volt ac power, and caution must be used when working near these circuits in the railroad signaling enclosure Often the traffic engineer will require an advance warning prior to the operation of the crossings warning devices. This may be as little as 5 seconds, or as much as 30 seconds in unusual situations. This advance warning allows time to alter the normal cycle of the traffic lights, and 4-125

Overview of Railroad Signal Circuits

can be used to flush out traffic traveling across the grade crossing. This is usually done by providing a green traffic light for several seconds to cars traveling across the grade crossing, followed by a yellow and a red light. Ideally, this sequence is completed before the crossings warning devices are activated, so that any accumulated highway traffic has been allowed to proceed, and the residual traffic has been brought to a stop before approaching the grade crossing. Sophisticated traffic preemption systems do all of the above, and then begin executing a modified traffic light routine, wherein all traffic, which does not need to travel across the grade crossing, is allowed to flow normally, and all other highway traffic is stopped. Alternatives to Track Circuits Although track circuits are the dominant standard for train detection in North America, the quest for improved signal system performance has led to the exploration of many alternative technologies. However, the successful implementation of new signaling and train control technologies lags behind the inventors in this area. For example, the use of slotted waveguides along the track to convey radar signals to moving trains was ready to go immediately following W.W.II. However, this technique still has not yet come into use. More recently, proposals have been advanced which call for optical fibers to be imbedded lengthwise in the railhead, or under the tracks. This would use Optical Time-Domain Reflectometry (O.T.D.R.) to observe the location of minute deformations in the cladding of the fiber, under the weight of each wheel. By doing so, the speed, direction, and exact position of a train could theoretically be measured in real-time. Another system that has recently been proposed is the use of trackside buried magnetometers to detect the presence of a train by its effect on the earths magnetic field (see below). By being aware of the myriad proposed solutions which have come and gone over the decades, persons responsible for solving railroad EMI/EMC problems can avoid investing time and resources in propositions that offer no realistic chance of successful outcome. With the possible exception of Robinsons original track circuit, progress in railroad signaling has generally been evolutionary in nature, rather than revolutionary, and it will likely remain so for the foreseeable future. Communication-Based Crossing Warning Systems One of the market forces driving invention in the area of crossing warning systems is the relatively wide gap between the cost of a simple crossbuck sign placed at a rural low-traffic grade crossing, and a conventional crossing warning system with lights, gates, bells, and a signal equipment enclosure with large backup batteries and an electric utility connection. In an attempt to fill this gap, a system consisting of two different types of solar-powered standalone flasher units with low-power radio transceivers and GPS receivers has been developed by

4-126

Overview of Railroad Signal Circuits

C3 Trans Systems LLC, and has been under evaluation for use at rural grade crossings previously equipped only with a crossbuck sign. The transceivers listen for a broadcast from a low-powered on-board radio beacon carried by the train. This beacon continuously transmits the trains position, as determined via an on-board GPS receiver (and other methods), and the flasher units activate themselves when the train is determined to be approximately 30 seconds away. Magnetometer and ultrasound technologies are used at the crossing to determine the occupancy status of the island. Bidirectional communications between the unit on the train and the crossing units provides additional system health monitoring and data recording capabilities. Although this system relies on non-vital radio communications in order to activate the crossing warning devices, which may prevent it from being considered vital in the strict railroad sense of that term, it may also hold significant promise for improving safety at rural grade crossings in a non-vital fashion. The potentially lower installed cost of this crossing warning system, as compared to conventional crossing warning systems, may make it especially attractive for installation at the large number of rural crossings in North America presently equipped with only a crossbuck sign for (passive) crossing protection. Magnetometers The distorting effects that iron and steel have on the Earths magnetic field can be detected by a device known as a magnetometer. Magnetometers have been used by the navies of the world to detect the presence of large pieces of ferrous metal for decades, perhaps most notably in their detection of submarines. As trains are made primarily of ferrous metal, this would seem to be an ideal application of this technology. One company, EVA Signal, of Omaha, Nebraska, has developed a crossing warning system based on this technology. However, at this time, these systems have been tested at limited locations, and have yet to be embraced by the railroad signal world as a whole. Wheel Detectors Wheel detectors are a form of presence detection that responds not to the presence of a shunt across the track, but the physical presence of a passing railroad wheel. These are often used in conjunction with wayside hazard detectors. These are similar to the wheel detectors used in Axle-Counter signaling systems. Some wheel detectors consist of a variable-reluctance magnetic device that works in a fashion similar to a magneto. The passing wheel flange changes the flux pattern surrounding a permanent magnet core, and the rapid changes in flux induce current into a coil that is wrapped around the core. These wheel detectors can be adversely influenced by heavy ac or transient dc currents flowing through the rail, as can happen on electrified railroads. More sophisticated wheel detectors 4-127

Overview of Railroad Signal Circuits

employ an eddy-killed oscillator circuit that is detuned by the presence of the wheel flange. This type of wheel detector is generally considered to be traction-current-immune. Hall-effect and Wiegand-wire technologies are also used in wheel detectors. Mechanical treadle wheel detectors are still common in Europe. Axle Counters The development and deployment of signaling technology in North America generally preceded corresponding work in Europe, Asia, and elsewhere. Because track circuits were immediately successful as a means of train detection, their use in Signal Control and Route Locking became a de-facto standard throughout the U.S.A. and Canada long before U.S. federal regulations enacted in 1939 mandated the use of track circuits, without exception, as the only permissible means of train detection. When later forms of train detection became available, most notably AxleCounting Block systems using the Wheel Detectors mentioned above, they were accepted outside of the U.S.A. and Canada in lieu of the track circuit. Outside of the U.S.A., axle counters continue to enjoy widespread and increasing usage on lines with light to moderate traffic. Axle-counting block systems have not gained acceptance in the U.S.A. because they are, at best, inferior to the track circuit in terms of functionality and performance. For example, axle counters are incapable of detecting a broken rail, which is one of the primary functions performed by track circuits. Axle counters do offer some benefits, particularly under circumstances where a track circuits shunting performance or low ballast-resistance capability may be in doubt, or where track circuits become unduly expensive to install and maintain due to track construction or electric propulsion considerations. Recently, the FRA granted permission to substitute axle-counting block systems for trouble-prone track circuits on a group of steel-deck bridges. In many instances, the axle counting system is less expensive than track circuits to install and maintain, particularly on light traffic single-track main lines. However, the onerous periodic inspection requirements attached to the approved FRA waiver have done little to encourage further consideration of this technology as a substitute for the track circuit in the United States. Global Positioning Systems When traditional track circuits seem unworkable, the now widespread use of the Global Positioning System, or G.P.S., is often the first solution that occurs to engineers unfamiliar with the intricacies of railroad signaling. Although it holds great promise for the future, and will likely be a component of many significant advances in railroad signaling, it suffers from several drawbacks as a sole means of train detection. The ARES system, developed and tested by Rockwell International and the Burlington Northern railroad in the 1980s, was a first attempt to build a signaling system using G.P.S. to determine the locations of trains. It suffered from several difficulties, the most serious of which was probably its positional accuracy. With closely-spaced parallel tracks, the system initially lacked the ability to determine exactly which track a train was on with absolute certainty, and so this system faced significant technical challenges from the beginning.

4-128

Overview of Railroad Signal Circuits

The recent peacetime removal of the Selective Availability encoding, which normally degrades the accuracy of commercial G.P.S. receivers, may allow this technology to develop further. Furthermore, the accuracy of properly equipped G.P.S. receivers can be enhanced by the reception of a reference signal from a local earth-based transmitter in a technique known as Differential G.P.S.. Such differential G.P.S. stations are becoming increasingly widespread in the U.S.A., thanks to active efforts by the U.S. Coast Guard (ships use them for precision navigation while docking in foggy ports), and the Federal Aviation Administration (F.A.A.), 25 which refer to this as W.A.A.S. . Any system proposing to use G.P.S. to replace track circuits for train location will face several engineering and economic challenges. First, it will have to provide such obvious cost savings to the railroad that its adoption will be an economic necessity. Second, it will need to be interoperable with the rest of the existing signal system used by a railroad, in order to allow for a transition period of perhaps several years. Thirdly, it must continue to provide all of the functionality currently provided by a track-circuit-based signal system. It is perhaps this third requirement that has prevented the adoption of the majority of new systems proposed. And finally, it will have to offer some advantage or benefit that is currently unavailable from the existing systems. To illustrate these requirements, lets look at an example: If you place a G.P.S. receiver on a locomotive, then the locomotive can continuously determine its location using the signals from the G.P.S. satellites. And the location of the rest of the cars in the train could be calculated, using the location and direction-of-travel of the locomotive, and a database of information including the consist of the train. But what if a single car got loose or was accidentally left behind during switching operations? In the 1990s, the brakes on a string of railroad cars, which were parked on a siding in the Los Angeles area, were apparently released by vandals. These cars began rolling downhill, and managed to enter a main line. The incline of the main line was favorable, and the cars continued to roll. After traveling for several miles completely unattended, they finally stopped when they collided at low speed with other cars on the track. No one was injured. A similar incident happened roughly a decade later, when the brakes on a string of cars were not properly set, and the unattended string of cars rolled for many miles before derailing. But there were no accidents at the grade crossings they passed through, nor were any other trains allowed by the signal system to enter the track they were on. Fortunately, track-circuit-based signal systems do not discriminate between intentional and accidental railroad traffic. And although many motorists were probably surprised and confused to see a string of cars moving on their own through a grade crossing, these motorists were never placed at risk. All of the crossing warning systems along the way functioned perfectly, providing the proper amount of warning time at each grade crossing. Being of modern design, the

25

W.A.A.S. stands for: Wide-Area Augmentation System. Eventually, a network of differential G.P.S. stations will be established at airports and other locations across the United States, making high-accuracy positional fixes possible almost anywhere within the continental U.S.

4-129

Overview of Railroad Signal Circuits

microprocessor-based crossing warning systems even recorded the speed of the cars, which was later downloaded during the accident investigation. In order to provide this sort of robust functionality and safety, a G.P.S.-based system could conceivably require the installation of a G.P.S. receiver and some form of 2-way radio communications equipment on every single railroad car. Maintenance and reliability concerns aside, the overwhelming cost implications of this solution are obvious. Hazard Detectors There is also a third category of equipment that we have assembled for purposes of this railroad signaling equipment overview that of Hazard Detectors. Hazard detectors are commonly utilized in both signaled and non-signaled territory to alert railroad personnel to the presence of unusual hazards that could affect the safety or efficiency of train operations. Hazard detectors may be classified in two distinct categories:
Roadway Detectors Rock Slide Earth Slippage Bridge Displacement High Water Fire (bridges) R.O.W. encroachment Vehicle Detectors Hotbox (overheated axle bearing) Excess Clearance Dragging Equipment Hot Wheel (sticking brakes) Broken Flange Loose Wheel Cracked Wheel Wheel Impact Load (flat spots on wheels) Dragging Brake Hose Couplings Track Scale A.E.I. Tag Reader

In the past, many hazard detectors such as dragging equipment detectors were interconnected with the signal system. When any hazard was detected, a restrictive signal indication (usually a red wayside signal meaning STOP) was presented to the affected train(s). The trains crew would stop the train, and investigate all potential hazards. With the advent of microprocessor-based talkers (hazard detectors that broadcast specific synthesized or pre-recorded warning messages directly to trains via VHF radio), now only railway hazard detectors such as slide fences remain interconnected with the signal system in the old way.

4-130

Overview of Railroad Signal Circuits

Wheel Defect Detectors There is an amazing and ever-expanding variety of devices designed to perform an automated inspection of the geometry and general physical condition of railroad wheels. Most of these are used to detect advancing failure conditions in the wheels themselves, such as chips, cracks, and flat-spots. However, with respect to the other types of equipment used on railroads, these are still relatively rare in North America, and will be found only in limited numbers on a given railroad. Hotbox Detectors In the case of hotbox detectors, a wheel detector or short track circuit controls operation of the hotbox detector so that the infrared sensor only scans its field-of-view when there is an axle bearing passing by. This ensures that the hotbox detector responds only to heat produced by a passing bearing. Otherwise, false positives could be reported by the system in response to direct or indirect sunlight, or some other unwanted source of infrared energy. This would require the train to stop, and the trains crew would have to physically inspect every axle of the train usually on foot.

Figure 4-84 Typical Wayside Hazard Detector Location having Hotbox and Hot Wheel Detectors

The delay caused by this false alarm may not affect only the stopped train, but other following trains as well. This can be very costly for a busy railroad line. Slide Fences Slide fences are systems designed to detect the displacement of nearby rocks, dirt, snow, or trees, which might foul (obstruct) the tracks and potentially cause a derailment. The simplest form of a slide fence consists of a long thin wire strung between poles where it would hopefully be broken

4-131

Overview of Railroad Signal Circuits

by a significant displacement. In the picture below, the slide fence is constructed on the poles overhanging the track.

Figure 4-85 Rock Slide in an Area Protected by an Overhead Slide Fence (Note the many insulators on the pole arms.)

Other forms of slide fences include boards or paddles, connected to a switch circuit controller, which can be knocked down or moved by an avalanche. Tag Readers Almost every railroad car in use now carries a small electronic device attached to its side known simply as a tag. This device is actually a transponder, which may be passive or active in nature. Passive tags can carry a limited amount of information, usually little more than a number which uniquely identifies the car. Active tags can have on-board memory, sensors, and battery back-up. In either case the concept is the same. Railroads have a need to know where every carload of goods is at any given time. By scanning the tag number of every car as it passes a location on the railroad, large railroads can easily track long-distance shipments for their customers.

4-132

Overview of Railroad Signal Circuits

But beyond simple carload tracking, an increasing number of on-board parameters are being measured. For example, the temperature of a refrigerated shipment can be monitored, and if dangerously high, the trains crew could be alerted to the problem. A.E.I. tag readers (for: Automatic Equipment Identification) may use wheel detectors to determine the trains direction of travel as the train moves past the tag readers interrogator antenna. Knowledge of the trains direction is necessary in order to make sure that the goods are heading in the proper direction. End-of-Train Monitoring (E.O.T.M.) E.O.T.M. telemetry devices were first introduced in the 1960s, and their use became widespread throughout the U.S.A. in the mid-1980s, allowing for the elimination of cabooses on most freight trains. Cabooses had been used as a location for the trains crew to monitor the air pressure in the trains brake pipe, and to operate the brakes from the rear of the train in an emergency. The primary purpose of the E.O.T.M. is to transmit brake pipe pressure data to the locomotive engineer, usually via a dedicated 452 or 457 MHz UHF radio channel. Transmissions are routinely made once every 30 seconds, and are made immediately upon a change in brake pipe pressure or other data. To ensure continuous communications with the locomotive, wayside radio repeaters have been found necessary in tunnels and on tangent track in exceptionally flat terrain. In order to avoid interference from other nearby E.O.T.M. units, each E.O.T.M. contains a unique digital I.D. code. The head-end receiver unit must be manually preset by the engineer to link-up with the E.O.T.M. at the rear of their particular train before departure. Todays E.O.T.M. includes a high visibility flashing red marker, illuminated only in darkness. Some E.O.T.M.s feature a motion sensing capability, allowing the engineer to promptly accelerate the train after observing that all slack in the train has been taken up, and the last car of the train has begun to move. E.O.T.M.s used in mountainous territory are now required to provide two-way communications capability. This works in conjunction with a remotely controlled air valve inside the E.O.T.M., which allows the engineer to initiate an emergency brake application from the rear end of the train. This could be a lifesaver in the event that the trains brake pipe, somehow became 26 obstructed . An obstructed brake pipe could prevent the brakes from being applied on any

26

The brakes of a train are released by pressurizing the brake pipe, and are applied by removing the pressure from this pipe (its really a flexible hose at each car-to-car coupling). This is controlled by a pneumatic logic unit on each railroad car. Surprisingly, the brakes are actually applied by the use of compressed air contained in one or more of the reservoirs carried on each railroad car. The pneumatic logic unit refills these reservoirs while the brake pipe is pressurized, and releases the brakes once the reservoirs are adequately pressurized (or if the reservoirs are completely empty). Any subsequent drop in brake pipe pressure causes the pneumatic logic to use air from the reservoirs to apply the brakes. And a subsequent increase in brake pipe pressure releases the brakes again. The amount of braking is controlled by the amount of pressure reduction in the brake pipe, with distinct normal service and emergency braking modes. Variable braking force is possible only in the normal mode. Thus the brake pipe both controls the operation of, and supplies compressed air to, the trains brakes.

4-133

Overview of Railroad Signal Circuits

railroad cars beyond the obstruction behind the train. In a worst-case scenario, the obstruction would occur somewhere close to the locomotive, leaving most of the trains cars without brakes.

Figure 4-86 Typical End-of-Train-Monitor (E.O.T.M.) (U.S.&S. Inc.)

Although various alternative power sources including fuel cells, air-driven turbines, solar cells, etc., have been tried on an experimental basis, most E.O.T.M.s are powered by an ordinary rechargeable battery. Other Miscellaneous Track-Mounted Signal Equipment Gauge Plates and Gauge Rods There are numerous locations along a railroad track, particularly in the area of switches, where the gauge of the track must be carefully maintained despite unusual lateral forces. In these instances, special structures are attached to the rails to keep them at the correct gauge (railhead-to-railhead spacing).

4-134

Overview of Railroad Signal Circuits

Figure 4-87 Gauge Plates and Gauge Rods

In order to be compatible with the signaling systems used, these structures must provide a mechanical connection between opposite rails of the track, without creating an electrical connection.

Figure 4-88 Insulation of Switch Gage Plates and Rods Prevents Unintended Short Circuiting of Track Circuits (U.S.&S. Inc.)

4-135

Overview of Railroad Signal Circuits

Gauge rods and plates are constructed using several of the same techniques as insulated joints, including bolts with insulating sleeves or thimbles on them, and insulating pads placed between the metal parts.

Figure 4-89 Deteriorated Insulation and Reduced Clearances on Gauge Plates

If the insulation on these structures should fail, the result may be a false shunt across the tracks. This false shunt may be present continuously, or may appear and disappear as a function of temperature. The actual impedance of this false shunt may vary widely, and may seem to have disappeared even though it is still present to some degree. This is possible due to the lowimpedance nature of railroad track circuits. A rough rule of thumb for signaling says that any shunt across the tracks, of an impedance greater than 10 Ohms, is essentially invisible to the signal system under most conditions. Switch Circuit Controllers Switch circuit controllers serve to detect the physical position of a track switch. They do this by having metallic contacts operated by cams that are indirectly mechanically coupled to the points of the moveable rails at a switch. Signal control circuits are passed through normally-closed contacts of the switch circuit controller. Signals governing train movement over the switch will display their most restrictive aspect if the points of the switch are open to a distance of either 1/4" or 3/8", depending on the type of switch (trailing or facing). Switch circuit controllers also may be used to detect the positions of other wayside devices such as: derails, slide fences, gates in fences leading into industrial plants, etc. 4-136

Overview of Railroad Signal Circuits

Figure 4-90 Typical Hand-Operated Switch with Circuit Controller (silver box in upper right corner)

Electric Switch Locks An electric lock is a device that locks a manually-operated track switch in place, when the operation of this switch could affect the safety of train operations. In general, electric locks are required on hand-operated switches in C.T.C. territory where trains are permitted to leave the main tracks via a spur track or un-signaled siding. This happens primarily when there are industrial spur tracks (sidings) that diverge from a main line controlled by a centralized traffic control system. The electric lock serves as a substitute for a remotely controlled interlocking control point at seldom-used hand-operated switches by preventing the switch from being opened (set reversed) when a conflicting train move is approaching on the main line. A time-release feature permits release of the lock once the padlock on the enclosure of the electric switch lock has been removed, and a pre-determined time interval has elapsed. A quick release feature is also provided to minimize the waiting time when exiting the main track, which is usually enabled by a simple overlay track circuit immediately adjacent to the switch. Electric locks within interlocking limits are controlled (released) by the dispatcher. Most electric locks are provided with an emergency release feature, allowing the train crew to break a seal on the electric lock enclosure, and release the lock if electrical power or communications between dispatch and the electric lock have failed. Two different types of electric locks have been installed in the U.S., and the selection of electric lock types is a matter of railroad preference: 4-137

Overview of Railroad Signal Circuits

1. lever-locking electric locks that serve only to lock the hand-throw lever of the switch stand, and 2. plunger-locking electric locks that provide an independent, redundant means of securing the switch points. Switch Machines Similar to its power-operated counterpart, the hand-operated switch machine has operating rod, lock rod, and point detector rod connections to the switch points.

Figure 4-91 Hand-Operated Switch Machine

Power-operated switch machines are usually associated with interlockings or C.T.C. control points where traffic conditions justify a means of quickly aligning routes prior to the arrival of a train. By aligning the switch to the correct track before the train arrives, this prevents the train crew from having to stop and manually throw the switch. Power switch machines are sometimes used on a stand-alone basis to save time in switching yards and other places where frequent manipulation of switches is required, and in some yard applications may be thrown by the train crew by remote control.

4-138

Overview of Railroad Signal Circuits

Figure 4-92 Typical Electric Yard Switch Machine with Circuit Controller (U.S.&S. Inc.)

A majority of power switch machines in service throughout North America are electricallycontrolled electrically-powered mechanisms employing a DC motor rated at 24 or 110 volts dc. DC power is often used, because of the need for battery backup. This motor is connected to a reduction gearbox, which provides the necessary mechanical advantage to move the rails or points of the switch. Many of these switch machines are equipped with a dual-control lever, enabling train crews to manually throw the switch in the event of a power or communications failure.

Figure 4-93 Typical Dual-Control Electric Switch Machine

4-139

Overview of Railroad Signal Circuits

Used primarily in large, busy passenger terminal interlockings, the electro-pneumatic (E-P) switch machine is electrically controlled and pneumatically operated. The simple, rugged E-P switch machine provides more rapid action than its all-electric counterpart. However, the expense of installing and maintaining an air distribution system has led to a gradual decline in the total number of E-P switching yards. Switch machines employing self-contained electro-hydraulic mechanisms of European design have been installed in a handful of locations throughout North America. In general, the electrohydraulic design has not proven advantageous over the simple, rugged, all-electric mechanism using a gear train for torque multiplication and a scotch yoke for rotary-to-linear motion translation. Similarly, linear motors have yet to be demonstrated as a practical means of operating switch points.

Figure 4-94 Typical Electro-Pneumatic Switch Machine and Valve

Figure 4-95 Electro-Hydraulic Switch Operator (RTI, Inc.)

4-140

Overview of Railroad Signal Circuits

However, in recent times, a substantial number of low-cost self-contained electro-hydraulic switch operator packages have been installed in small switching yards, and on non-signaled secondary trackage. Using solar energy and a photovoltaic cell to recharge a 12-Volt battery, the electro-hydraulic switch operator serves to reduce the physical burden associated with the frequent throwing of hand-operated switches in non-signaled territory. Most are pushbutton operated, however, some switches of this type have been equipped with radio-based remote controls. Some switch machines used for trolleys and other street railways employ 600 Volt dc solenoids as their drive mechanism, due to the availability of this type of power on those systems. Electric switch machines of European or Asian manufacture for trolleys usually employ 380 V ac or 600 V ac motors. In terms of susceptibility to ac interference, switch machines themselves are generally not susceptible, but the circuits controlling them can be. Impedance Bonds Although they were briefly discussed together (above), the impedance bonds used in dc propulsion territory differ significantly from those used in ac propulsion territory. This is due to the differences in the nature of the propulsion current. DC systems often use low (600 Volt) propulsion current delivered via a third-rail conductor. This contrasts with ac systems, which often use voltages approaching 25 KV from an overhead wire. In order to deliver the same amount of power, the currents involved in these two systems differ greatly. One function of an impedance bond in dc propulsion territory is to provide the highest possible rail-to-rail impedance for ac signaling frequencies. In order to achieve this, a large laminated magnetic core must be wound with as many turns as possible. As the propulsion return currents are generally much higher in low-voltage (600 Volt) dc propulsion territory, only a limited number of these necessarily thicker windings can be fitted into the allotted space. For power-frequency track circuits in dc propulsion territory, a typical impedance bond with a continuous 2500 amp-per-rail rating weighs roughly 1000 lbs, contain 17 gallons of mineral oil (for heat dissipation), and has a 10-turn track winding composed of copper strips with a crosssection measuring 1/4"x5". Its rail-to-rail dc resistance is less than 30 micro-ohms. Higher rail-to-rail signaling frequency impedances and higher ampacity ratings could be readily obtainable. However, such a bond would probably be too big to fit between the rails, which is where impedance bonds have historically been located. This is not necessarily a problem, since many railways now place their impedance bonds off to one side of the track, away from the potentially damaging reach of mechanized track-maintenance equipment. In some cases, the use of concrete ties or slab track construction precludes the placement of impedance bonds between the rails.

4-141

Overview of Railroad Signal Circuits

Figure 4-96 Typical Construction of Impedance Bond Used on Power Frequency Track Circuits in DC Propulsion Territory

Ideally, opposing fluxes produced by equal currents flowing through the side leads of an impedance bond will cancel, producing little or no net flux within the core. In real-world operation, however, unbalanced flow of DC current flowing through opposing halves will significantly magnetize and saturate the core. A significant degree of saturation in the impedance bonds core is detrimental because it reduces the bonds rail-to-rail ac impedance. Under severe unbalance conditions (greater than 10-12% of the net ampacity rating) the rail-to-rail impedance drops to nearly the dc resistance of the windings, and the track circuit becomes shunted by the impedance bond. Once the unbalanced current flow ceases, the rail-to-rail impedance returns to its nominal value. To reduce the high levels of flux produced by imbalance, the core is provided with an air gap, typically between 0.040" and 0.170" in length. The air gap feature is an undesirable compromise, however, since it significantly reduces the nominal rail-to-rail impedance. To counteract the adverse effect of widening the air gap, the impedance bond may be resonated (parallel resonance looks more like an open), boosting its rail-to-rail impedance from 0.4 to 2.0 ohms at the signaling frequency (typical values for a specific type of bond). Where multiple track circuit frequencies are used, the bond may be made resonant at more than one frequency.

4-142

Overview of Railroad Signal Circuits

Figure 4-97 Typical Impedance Bond Layout, 750 VDC Propulsion Territory, 100 Hz Track Circuits; Impedance Bonds Rated 2500 Amps Per Rail

Aside from improving the impedance bonds ability to tolerate imbalance, the resonating of impedance bonds is often beneficial in improving track circuit performance, particularly in cab signal territory where a very high feed-end voltage (>18 volts) is required to provide 3 amps of axle current at the opposite end of a 4,500-foot track circuit. With un-resonated bonds, much of the power would be dissipated (wasted) in the feed-end bond. Also, the cab signal transmit current would be unacceptably high, perhaps over 20 amps at the feed-end of the circuit. One difficulty with resonated bonds is the incompatibility between resonated bonds and AFOs. Unless the resonating circuit is specifically designed to accommodate AFOs, the bond will generally exhibit a low rail-to-rail impedance at frequencies below 10 KHz. Although some very old impedance bonds have rail-to-rail impedances as low as 0.25 ohms at 25 Hz, these are generally limited to applications on very short track circuits with low (0.06 ohm) shunting sensitivity. The added cost and diminished reliability of the resonating winding, capacitor, and surge protector are significant considerations. Oil-filled resonating capacitors rated at 660 V ac are susceptible to failure when subjected to very high voltage spikes, produced by a sharp asymmetrical step change in the dc current flowing through the bond, as may happen when the first or last axle passes over staggered insulated joints.

4-143

Overview of Railroad Signal Circuits

Figure 4-98 Various Types of Resonating Circuits for Power Frequency Impedance Bonds

One method to improve this situation is to add surge protection devices across the resonating capacitor. However, surge protectors tend to have a finite life, resulting in a detuning of the resonating circuit when the surge arrestor becomes shorted. Alternatively, this may result in a failure of the capacitor after the suppressor has become ineffective in suppressing surges. Surge arrestors so used must be protected against contamination by oil or moisture. In one incident, a high current fault between the third rail and running rail destroyed the resonating units in six nearby impedance bonds. Impedance bonds manufactured per AREMA recommendations must maintain at least 90% of their nominal rail-rail impedance when subjected to a 12% imbalance condition (i.e. 180 amps for a 1500 amp-per-rail bond). Many impedance bonds manufactured outside the U.S.A. have a relatively high rail-to-rail impedance, but low imbalance tolerance. The size, weight, and manufacturing cost of an impedance bond is primarily dependent on its ampacity rating. Differing techniques for testing and interpreting test results have been developed over many years, sometimes yielding inconsistent or incomparable ratings. Because impedance bonds are expensive (up to several percentage points of the entire capital cost of railroad electrification), their characteristics and ratings are frequently in dispute.

4-144

Overview of Railroad Signal Circuits

Figure 4-99 Typical Impedance Bond Layout, 12 kV, 25 Hz Propulsion Territory, 100 Hz Track Circuits; Impedance Bonds Rated 300 Amps per Rail

Figure 4-100 Typical Impedance Bond Layout for Audio-Frequency Track Circuits, 750 VDC Propulsion Territory, 1-5 kHz Track Circuits; Impedance Bond Rated 3000 Amps per Rail

4-145

Overview of Railroad Signal Circuits

Other Signaling Topics Electric Traction Circuits On an electrically propelled railway, traction power is distributed to moving trains via an overhead wire or third rail. Power is generated centrally, and transmitted to railroad-owned substations located every 1-25 miles along the railroad line. Electric traction systems invariably dictate the use of more sophisticated signaling systems, because: 1) the signal system must be absolutely immune to the effects of stray propulsion currents, and 2) electric traction systems are typically used only in dense traffic corridors where high speeds and short headways between trains are the norm. Issues involving the inter-relationship of signaling and electric traction systems are pertinent to the subject of utility-railroad EMC coordination. During the 1890s, various forms of dc electric propulsion systems for railways were developed; the first ac electrification occurred circa 1911. During this era, electrification as a substitute for steam motive power overcame many obstacles and inefficiencies in railway operation. Electric propulsion of trains was recognized as mandatory for underground rapid transit systems. Elsewhere, electric propulsion allowed economical, one-man operation of single-car trains such as urban and inter-urban trolleys. Rapid acceleration, bi-directional operating capability and high rolling stock availability prompted electrification of many busy commuter districts. Some railway electrifications exploited inexpensive hydro-electric energy resources. Electric propulsion in mountainous territory was advantageous for several reasons, most notability the ability to temporarily overload the electric locomotive to nearly double its continuous horsepower rating, in some instances for up to an hour or more. The benefits of dynamic and/or regenerative braking also were important considerations where loaded coal trains descended long grades. Worldwide, electric propulsion of trains is widespread and slowly growing. For new mainline and high-speed rail systems not required to be interoperable with existing wayside equipment and rolling stock, the de-facto standard is 25 kV, 50 Hz. Many older systems employing 3 kV dc or 15 kV, 16-2/3 Hz will remain in-service indefinitely. Systems operating at 50 kV, 50 Hz have been built. However, the savings made possible by the increased substation spacing at this voltage do not appear to offset the added cost and other disadvantages of high catenary voltages. With the exception of small people movers, rapid transit systems such as subways almost invariably employ low voltage (600 volt) dc energy distributed by third rails. As with other types of electric traction systems, the track (one or both running rails) serves as the return path. One notable exception is the extensive 4-rail system of the London Underground, where separate positive and negative contact rails are provided. In this system, the running rails are electrically isolated from the propulsion system. The most important advantages of third rail systems over catenary-wire systems are the reduced vertical height of the tunnels required, and the elimination of the complexities surrounding the installation of catenary supports on bridge structures. In the U.S.A., nominal third-rail propulsion voltages range from 550 to 1000 V dc. Elsewhere, third rail voltages up to 1500 V dc have been used.

4-146

Overview of Railroad Signal Circuits

Unavoidable reactance losses in the third rail and track eliminate any consideration of ac power with third-rail systems, since very high currents are needed to develop several thousand horsepower at 600-750 volts. Additionally, third rail voltages cannot be raised significantly without encountering significant difficulties with clearances, flashover, and personnel safety. Distribution of the motive power throughout the multiple-unit train (typically 500-750 horsepower/car) has also proven advantageous because the propulsion current is collected by a large number of current-collection shoes. In a dc propulsion system, the insulated contact wire or rail is always of positive polarity with respect to the un-insulated running rails that form the return path. This approach reduces, but does not eliminate, the possibility of electrolysis in an underground metallic object in the vicinity of the railway and its substations. Cases on record show that corrosion and rupture of newly-laid underground pipelines can occur in as little as 12-18 months as a result of electrolytic corrosion from incorrectly-polarized dc propulsion currents. Lead-sheathed cable, the steel bars used to reinforce concrete, and metallic tunnel shells are also subject to electrolytic wasting. Electrolysis cannot occur unless the underground conductor has a positive potential of about 0.75 volts or higher with respect to the surrounding earth. In some instances, impressed potential systems have been installed to counteract both natural and manmade dc currents; these systems pass current from a sacrificial positive anode to maintain a negative potential on the protected object, with respect to remote earth. Electrolysis is a pervasive problem usually mitigated by careful maintenance of rail bonding. In some instances, the resistance of the return network must be reduced by augmenting its conductivity, typically by using a 2,000 kcmil cable paralleling each track. Embedded wires, screens or reinforcing bars are sometimes employed to intercept dc current leaking from the rails to earth. The return network for yard trackage, which is typically of low ballast resistance, may be isolated from mainline trackage by contactors or diodes to minimize sinking of dc propulsion current into the earth. In rare instances, contactors may be installed to isolate certain sections of track unless or until they are occupied by a train. Electrolysis problems also are mitigated by modern track construction and maintenance techniques that exhibit very high rail-earth resistance. Close cooperation between the railway and affected utilities is essential to controlling electrolysis. Long-standing electrolysis control committees have become an institution in some major cities.

Figure 4-101 Electrolysis of Underground Structures Caused by Unwanted DC Current Flow

4-147

Overview of Railroad Signal Circuits

Unwanted dc current flowing through the electric utilities multi-grounded neutral networks is generally attributable to the same causes as electrolysis, namely poor bonding or other inadequacies in the railways propulsion current return network. Very high levels of stray dc may tend to open fuses or damage conductors and equipment such as transformers. Lesser amounts of stray dc current flow produce more subtle effects, including electrolysis of ground rods and water pipes at negative potential points where dc current leaves the conductors. Another serious problem involving stray dc current is the asymmetrical saturation of transformers. The dc current flowing through the transformers primary or secondary winding produces a non-varying flux in the core. As the ac flux component reaches its peak in the same direction, the core material is driven into saturation. The result is transformer overheating and waveform distortion with strong 4th-harmonic content. Similar difficulties due to core saturation are sometimes experienced on long-distance transmission lines due to geo-magnetically induced currents attributable to solar flares. This problem is particularly difficult to investigate due to the wide fluctuations in localized dc current flow associated with moving trains. AC propulsion systems have several very important advantages over dc systems: Freedom from the scourge of electrolysis; Less copper and lighter weight of catenary; Greater spacing between substations; Ability to deliver higher levels of power at high speeds; Ability of protective relaying system to discriminate between loads and faults; Lower, controllable fault currents;

Figure 4-102 Electric Traction Fed by Overhead Wire and Contact Rail

4-148

Overview of Railroad Signal Circuits

Figure 4-103 Impedance Bonds with Electric Traction

Figure 4-104 Various Wayside Signal Mountings

Safety, Vitality, and Fail-Safe Design Concerning matters of safety and efficiency, the importance of railway signaling is underscored by the fact that more than half of the annual 1.5 trillion ton-miles of traffic in the U.S.A. is carried over only 10% of its railroad network. High-density freight corridors typically handle in excess of 150 million gross tons per year on two or more main tracks. Freight train movements cannot be efficiently orchestrated in accordance with rigid schedules. Traffic flows tend to be sporadic with pronounced bursts and lulls. Railroads have little control over the seemingly limitless number of customer-related factors that may affect the loading, movement, or unloading of cars. Efficient signaling substantially increases the productive capacity of railroads and their finite, costly assets such as track, motive power, rolling stock, train crews, etc. The importance of achieving safety in control is illustrated by a series of 12,000-ton trains passing at 50 M.P.H. several times per hour on a busy freight line. 4-149

Overview of Railroad Signal Circuits

Any train can be operated in any territory without the benefit of a signal system, but only at a substantial penalty in terms of increased operating expense, greater transit time, and diminished productivity. Because persons involved in railway signaling matters are generally intelligent, career-conscious engineers and craftsmen who must accept personal responsibility for the safety of the signaling system, signaling technology continues to evolve slowly and cautiously. Many of todays design principles, equipment, and materials would be familiar to signal engineers from the 1920-1960 era. The overall architecture of a railway signal system may be described as non-redundant, singlefault-tolerant, and ultimately dependent on devices, circuits, and system having fail-safe characteristics. Railroad signal systems are designed and constructed, insofar as possible, such that any disarranged or broken part will cause one or more signals to display their most restrictive indication. Under these conditions, train movements will either be stopped, or will be prevented from starting. With few exceptions, virtually all railway operators worldwide employ fail-safe signaling systems using specialized components and design techniques. Some European authorities have dismissed the notion of fail-safe components, opting instead for equivalent protection by means of cross-checked redundant logic elements. In the U.S.A., the use of specially designed fail-safe signaling components in vital applications has proven highly successful over a period of many decades. Although train wrecks are significantly more frequent in certain third-world countries, they are generally not attributable to signal system failure. Rather, the vast majority of these railway accidents are attributable to just two root causes: unsafe operating practices, and equipment defects. While increasing emphasis has been placed on exhaustively investigating the circumstances and cause of each railway accident in recent years, few, if any, fundamentally new causes have appeared over the past 100 years or so. The term vital, as used by railroad signaling engineers, carries the connotation of being essential to safety. In this sense, the term vital is synonymous with seaworthy or airworthy in the marine and aviation communities, respectively. Corresponding nuclear power system components critical to safety would be termed Class 1E. Non-vital components are those components whose failure or mis-operation cannot adversely impact the safety of train movements. The terms vital and fail-safe are often used interchangeably and indiscriminately, yet neither term is explicitly defined under current FRA rules in 49 CFR 234 or 236. Another key principle in the design of railroad signaling equipment and systems is the requirement that the potential vital failure modes of a device or system must be self-revealing, and show themselves in the form of a restrictive signal indication, or the inability to throw a switch. Periodic inspection and testing of railroad signal equipment is required only where electrical or mechanical faults are not reliably self-revealing. Remote monitoring and diagnostic functions have not yet been accepted as a substitute for fail-safe design, or periodic test and inspection. 4-150

Overview of Railroad Signal Circuits

Federal law in the U.S.A. requires that the cause of each signal failure must be determined, and the underlying fault condition corrected without undue delay - which usually means prior to the next train movement. Every False Proceed or Potential False Proceed failure must be reported to the FRA within twenty-four hours of its detection. An obscure but invaluable feature of todays signaling technology is its frequent ability to conclusively identify the specific component (or person) responsible for each and every failure. Correction of a defect thus requires only identification and replacement of the defective component or errant individual. Agonizing guesswork, endless testing, and tenuous risk analysis are unnecessary except in the most extreme cases. Unfortunately, EMC problems generally cannot be easily identified or resolved using simple on/off test and substitution techniques. Despite the many millions of pieces of railroad signal equipment in service today, actual falseproceed signal indications are a fairly rare event. Approximately 150-200 false-proceed and potential false-proceed conditions are reported to the FRA annually, not all of which can be confirmed. Accidents caused by the failure of signal equipment occur at a rate so low as to be statistically insignificant, meaning that no discernable trends can be found in analyzing this data. The failure modes and mechanisms are well understood and effectively controlled by FRAmandated test and inspection requirements. The very nature of this business has dictated that U.S. railway signal equipment manufacturers maintain a policy of conservative design practices and rigorous quality control. Many signal appliances manufactured today are direct descendants of designs initially created over 75 years ago, and many signal appliances manufactured 50 years ago are still reliable and in service today. The design principles and safety requisites for railway signaling systems long pre-date MIL-STD-882 and related system safety analysis methodologies established in connection with U.S. military and space programs in the mid-1950s. Railway signaling stands as one of the earliest examples of a life-safety system where human life and safety are solely dependent on the correct functioning of man-made instruments.

Figure 4-105 Track Circuits Provide Broken Rail Detection Capability. However, as Shown by this Cracked Joint Bar Bridged with a Bondwire, Certain Defects are Undetectable in the Track Circuit

4-151

Overview of Railroad Signal Circuits

The signal design techniques that have evolved capitalize on certain unique features of the railroad. For example, the track circuit is a simple, highly reliable means of detecting train position. Nothing resembling to the track circuit has been developed for any other mode of transportation. However, track circuits will probably never be the perfect solution to all railroad signaling problems. The railroad is also unique in comparison to aviation or nuclear energy, because its trains can usually be safely stopped by passive default anywhere, anytime, in the event of a fault without incurring further risk. Despite continual economic pressure to minimize installation and maintenance costs, concepts such as statistical risk analysis, redundancy, or defense-in-depth have yet to be accepted as a substitute for time-tested fail-safe design principles. Because of their specialized design and construction features, signal system components have exceptional reliability and longevity. Signal systems and their components are usually among the oldest useful assets of any railroad, often fully depreciated for tax purposes in less than half of their useful life. Thirty to fifty year old signal installations are common today, and the ultimate life of a signal system is usually limited only by the integrity of its wire and cable insulation. Despite the enviable safety record that railway-signaling systems continue to enjoy, many important improvements in equipment design and maintenance practices have been made over the past 125 years. Often, the necessity of these improvements became evident only after some unfortunate incident involving property damage or loss of life. Because most such incidents were, in retrospect, obvious or predictable to cognizant personnel based on past experience, there remains a strong aversion to any innovation into uncharted territory, at least until more of the known problems with the existing signal technology have been successfully addressed.

Figure 4-106 Although Signal Systems and Track Circuits are Based Upon Fail-safe Design Principles, No System Offers Protection Against All Conceivable Hazards

4-152

Overview of Railroad Signal Circuits

The hazards arising from EMC issues are especially worrisome to the signal engineer, because the cause and effect relationship is often transient and thus difficult to observe and analyze. Fortunately, the fail-safe nature of the equipment assures that the vast majority of failures will be so-called right-side failures, in which the signal system takes appropriate action to protect the safety of the public. Rules, Standards and Instructions The technical requirements concerning railway signal systems in the U.S. are found in the Code of Federal Regulations (CFR), Title 49, Part 236. The corresponding regulations for grade crossing signal system safety are in 49 CFR, Part 234. Part 236 contains more than 150 individual Rules, Standards and Instructions (RS&I) pertaining to signals and train control systems. These regulations apply only to the operators of railroad signal systems; no technical or procedural regulations have yet been imposed upon manufacturers, distributors or installers of such equipment. However, railroads are quick to recognize equipment which does not assist them in their efforts to comply with these rules. The RS&I rules were enacted on September First, 1939, following passage of the Signal Inspection Act in 1937. Specific requirements and the text of the RS&I have been negotiated and re-negotiated on an ex-parte basis between Surface Transportation Board (S.T.B., formerly called the Interstate Commerce Commission or I.C.C.) or FRA staff members, railroads, and labor representatives. Signal equipment suppliers have generally avoided interjecting their voice in rulemaking matters, except at the specific invitation of their railroad clients. Important revisions to the RS&I were made in 1950, after it became apparent that certain rules concerning maintenance (i.e. the painting of signals) or operating efficiency fell outside the safety mandate of the Signal Inspection Act. Several minor changes were made between 1960 and 1980, followed by another major overhaul in 1984, and the addition of rules for microprocessor-based systems in 2005. Despite significant technological advancements over a period of 60+ years, the majority of rules in todays RS&I remain virtually unchanged from those established 1939. The RS&I cover only the basic technical requirements and is notoriously vague in many instances. Railroads and labor have long bemoaned the ambiguity of certain rules, along with the FRAs sometimes-inconsistent interpretation and enforcement actions. Nonetheless, establishment of one national standard is still far preferable to the chaos that would surely accompany the establishment of individual state codes in this area. As a bonus, the 1937 Signal Inspection Act also bestowed upon respondent railroads a clause precluding tort action based on allegations that a specific type of signal system might have prevented an accident. This means that the railroads could not be held liable after an accident just because they had not equipped every single wayside signal and grade crossing location with the very latest equipment. The FRAs jurisdiction is limited to mainline railroads and certain other rail carriers in the U.S. Most urban transit and light rail systems remain exempt from FRA regulations, because their tracks are not considered to be connected to the North American rail network. Several 4-153

Overview of Railroad Signal Circuits

paradoxes in the scope of FRA jurisdiction can be explained by the long-term political discourse that has taken place among the railroad owners, labor interests, local transit subsidy sources, and Congress. And even though they are not required to comply with FRA rules, in virtually all cases, railroads exempt from these rules have imposed contractual requirements for new signaling systems that dictate conformity to the applicable portions of 49 CFR 234 and 236. From a worldwide perspective, the RS&I rules establish what is arguably the highest attainable standard of safety while imposing only a very small burden on the overall cost of rail transportation. Regulations pertaining to EMI/EMC issues are completely absent from past and present editions of the RS&I. Since false proceed conditions stemming from power line interference are generally rare, and are always triggered by a unique set of circumstances, it appears unlikely that a regulatory solution could be imposed to prevent future incidents of this sort. Nationwide, the FRA employs a staff of about 50 signal and train control inspectors and specialists; none of whom are known to have been specifically trained or equipped to detect and analyze EMC conditions. Although FRA does not maintain specific rules covering EMI/EMC issues, it may issue emergency orders to restrict or suspend railroad operations until imminent safety hazards are corrected or mitigated to its satisfaction. Additionally, substantial civil and criminal penalties can be assessed against the railroad and/or individual railroad employees where there has been willful violation of FRA rules or orders. This usually ensures compliance with its rules. Lightning Protection and Grounding The field of railroad lightning protection and grounding is fraught with pitfalls in terminology. This is because many terms loosely used by railroad personnel have very specific and different meanings in the electric utility world. In general, the equipment used by railroads to protect their signaling equipment from unusual voltages on the rails is designed and intended to address issues arising from lightning strikes.

4-154

Overview of Railroad Signal Circuits

Figure 4-107 Typical Lightning Arrestors. Red Arrestors are Line-to-Ground, Blue Arrestors are Line-to-Line

Faced with repeated failures and equipment damage in high isokeraunic areas, railroads have long searched for tools and techniques that could reduce their susceptibility to lightning damage. The widespread substitution of solid-state components in place of relays and other electromechanical elements has resulted in an increasing frequency of failures caused by lightning, electrostatic discharge, and other types of surges. The effectiveness of surge protection devices and techniques remains unsatisfactory in many instances. The grounding of signal equipment is a complex, multi-dimensional issue. Although the requirements and benefits of grounding electrical systems are generally well established, signal control circuits are customarily and intentionally ungrounded. This is due to the vital nature of signal circuits. Multiple grounds can bypass open contacts in the signal system, falsely energizing a circuit. Similarly, multiple grounds on a circuit may provide a source of energy where differences in earth potential exist. By maintaining isolation from ground, any ground that develops is readily detectable, and has no adverse effect on the safety of the signal system until a second ground develops. Improper grounds are usually attributable to deteriorated insulation, mechanical damage to wire or cable, crosses between conductors and shorted lightning arrestors.

4-155

Overview of Railroad Signal Circuits

Figure 4-108 Entrance Terminals with Wide, Flat Grounding Strips

Figure 4-109 Effect of Multiple Grounds on Track Circuits

4-156

Overview of Railroad Signal Circuits

Recently, the basic principles of E.M. shielding have been applied in an attempt to effectively separate the dirty wiring entering the signal equipment enclosure from the clean wiring within the enclosure. The long-term effectiveness of these attempts has yet to be demonstrated, but the results of laboratory studies conducted to date are encouraging. AC Interference Issues As the two most sensitive and precise types of track circuits currently in use on North American railroads, Motion Sensors and Crossing Predictors are often the first to notice the presence of an ac interference problem on a railroad. As such, they are often blamed for the problem, even before a meaningful investigation into the cause of the symptoms has been conducted. This is a form of shooting the messenger. Railroads and power lines often share a right-of-way out of economic necessity. Often, the most economical land that can be obtained for the construction of new power lines is an existing railroad right-of-way. Considering how many miles of railroad trackage parallel ac power transmission and distribution lines, the frequency of conflicts between the power lines and railroad signal systems is actually quite low. The variety of exact symptoms that ac interference can induce in Motion Sensors and Crossing predictors is dizzying. But they all boil down to one of the two basic failure modes of these types of equipment. Either: a. The crossings warning devices are activated when they shouldnt be, or b. The crossings warning devices are not activated when they should be. Fortunately, the first circumstance, a right-side failure, is far more common than the second (i.e., a wrong-side failure). Furthermore, the necessary combination of circumstances required for ac power lines to bother a railroad signal system is usually very simple. In nearly every case, all of the following items must be present: 1. A source of electrical energy, such as the power lines 2. A path for the electrical energy to follow, either inductive (magnetic) coupling, capacitive (electrostatic) coupling, or a conductive path 3. An unbalance which causes the ac potential to be greater on one rail of the track than the other 4. A receptor (such as a Motion Sensor or Crossing Predictor) that is susceptible to the type of ac voltage present between the rails of the track. Eliminating or adequately mitigating one or more of the four items above will result in a return to normal signal system operation. This statement holds true for almost every type of track circuit used today. 4-157

Overview of Railroad Signal Circuits

Power lines are such a part of our modern lives that they can now be found at almost any point along a railroad line. This is what makes the mitigation of the source (1, above) perhaps the most difficult and expensive option when ac interference is encountered with railroad signaling circuits. Any metallic conductor, whether insulated or not, which is subjected to a time-varying magnetic field from some external source offers the possibility of elevated potential and undesired current flow due to magnetically induced energy. This so-called Longitudinal Induction is often commonplace on railroads. However, it is rarely troublesome, as it affects both rails of the tracks equally. Electrostatic induction is similar, in that the presence of a capacitive coupling path from the energy source to the rails will generally affect both of the rails equally. In general, the shorter the distance between the source of energy and the circuit being unintentionally coupled to it, the greater the induction.

Figure 4-110 Adequate Clearance Between Signal Equipment and Overhead Lines Must be Maintained

4-158

Overview of Railroad Signal Circuits

As almost all modern track circuits view the track differentially, i.e. they can only sense the difference in voltage from one rail to the other, the elevation in voltage of both rails that longitudinal induction normally causes is essentially invisible to the signal system. This means that the induction, although present, has no effect on the operation of the signal system. However, when there is an electrical imbalance present in the track, which causes one rail to be affected by the magnetic or electrostatic induction to a greater degree than the other, we now have one of the necessary ingredients for ac interference with railroad signaling systems. The presence of an imbalance is the catalyst that can convert common-mode voltages (both rails equally energized) into differential-mode voltages (rails unequally energized). Conductive paths, which provide a way for energy from any part of an ac power system to reach the railroad rails, can produce the same end results as the combination of induction and imbalance. That is, one rail of the track may be at an elevated ac potential with respect to the other. Most often, this occurs because of an unintended connection that has been made between the ac power system, and one rail of the railroad tracks.

Figure 4-111 Foreign Energy on Track Circuits or other Signal Facilities may appear when Electric Switch Heaters Become Grounded to the Running Rails

The last key ingredient is the piece of signaling equipment that acts as the receptor for the ac energy. Some types of equipment are just inherently more susceptible than others, because of their design, construction, and the laws of physics. But Motion Sensor and Crossing Predictor track circuits are not the only types of equipment on a railroad signal system subject to EMC/EMI concerns. The four essential elements given above for track circuits apply generally across the field of railroad signaling. AC interference is often a diagnosis reached incorrectly and all-too-soon by well-meaning signal maintenance personnel. It stems from their frustration and inability to find the actual underlying cause of the problem. Instead, ac interference should be a diagnosis that is reached only after the 4-159

Overview of Railroad Signal Circuits

discovery of direct evidence to support it, or after all other more-likely causes of similar trouble have been eliminated. AC interference is an appealing scapegoat, due to its variable nature. A symptom which does not occur consistently can plausibly be blamed on ac interference, which is known to wax and wane as a function of line current, power flow, phase imbalance, and many other factors. But this is not science. Nor does it move us in the direction of a solution. Only by understanding the nature of the equipment used can we hope to work towards the resolution of an Electromagnetic Compatibility problem involving ac power lines and railroad signaling equipment. Hopefully this chapter has moved its readers in that direction.

References
30. AREMA Signal Manual Part 8.1.5, 1999. 31. Batteries and Energy Systems by Charles L. Mantell, published in 1970 by McGraw-Hill, Inc. Library of Congress Catalog Card Number 70-107448.

4-160

5
ABNORMAL OPERATION OF RAILROAD EQUIPMENT
This chapter describes abnormal operation of railroad equipment, with specific concentration on mis-operation that can result from electromagnetic interference. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA. Bennett R. Feely has been engaged in the railway signaling profession throughout his career, beginning in Philadelphia, PA, as a signal maintainer on the subway-elevated lines of the Southeastern Pennsylvania Transportation Authority. He has since worked as a signal engineer for General Railway Signal and Union Switch & Signal throughout the past 25 years, specializing in wayside devices, circuits and systems. Areas of special interest include track circuits, switch machines, relays and light signals. Through his work and studies he has gained detailed insight and expertise in technical and commercial matters pertaining to the design, manufacture, installation and maintenance of railway signaling equipment. Mr. Feelys broad range of experience also includes participation in important projects such as NYCTs Canarsie CBTC project and Amtraks Northeast Corridor Improvement project. Mr. Feely attended the University of Pittsburgh and Penn Technical Institute. He holds several patents in the field of railway signaling.

5-1

Abnormal Operation of Railroad Equipment

Introduction
Abnormal operation of railroad equipment is one possible outcome of ac interference. When abnormal operation is caused by ac interference, it is usually some form of steady-state interference, because railroad signal systems are designed to deal with the very short-lived disturbances caused by faults and switching surges. Additionally, it is usually rail-to-rail (differential mode) interference that affects operation. This chapter includes a quick-start version. However, it is very highly recommended that every reader take the time to read the detailed version. While power engineers expect this to be new to them, this information will also be new to many railroad personnel as well. Training in railroad signal systems usually includes little study (if any) into the operational effects of ac interference.

Quick-Start Version
Railroad equipment is designed to fail in such a way that safety is maintained. So, if the track signals detect a problem, they slow or stop the trains. If the highway crossing gate system detects an inappropriate input, it lowers the gates. The idea is that it is better to stop people and freight, than to risk a collision. These safe failures are sometimes called right-side failures. The opposite would be wrong-side failures. Wrong-side failures are simply unacceptable. Nearly all abnormal operations of railroad equipment resulting from ac interference are right-side failures. However, we must always be alert to the possibility of wrong-side failures, and do everything possible to prevent them. The most common abnormal operations resulting from ac interference are false activation of highway grade crossing train detection equipment (the gates are down with no train), and dropout of 60 Hz cab signal systems. Because these systems operate at audio frequencies near power line frequencies and at very low power levels, they are more susceptible to ac interference than other systems.

Detailed Version
Abnormal Operation of Railroad Signal Equipment This chapter is devoted to the discussion of several engineering philosophy issues that underlie most discussions of railroad signaling systems and ac interference. Our objective here is to clarify the terms used, and discuss them without pretending to offer any quick fixes for the problems themselves. This chapter is recommended reading for anyone who is less than completely familiar with the web of issues faced by railroad signal equipment designers and manufacturers, and by railroad signal departments.

5-2

Abnormal Operation of Railroad Equipment

Definitions, Terminology, and Philosophy The term abnormal operation, as it will be used within this chapter, has a specific meaning that we will define. But in order to understand this term in the proper context, we must first explain the use of several other commonly used terms. Most discussions of ac interference that take place between power systems and railroad signaling personnel suffer from a lack of common terminology. The independent evolution of power systems and railroad signaling has resulted in each field developing its own unique dialect of the language of electrical engineering. Terms used in one field are often either unused, or have a different meaning, in the other. In particular, the terms vital and fail-safe in railroad signaling have acquired unique and specific meanings insofar as they are commonly used within the railroad signal world. Part of the Websters Encyclopedic Unabridged Dictionary definition for the term fail-safe includes the following: Fail-Safe 1. Electronics. pertaining to or noting a mechanism built into a system, as in an early warning system or a nuclear reactor, for insuring safety should the system fail to operate properly. 2. equipped with a secondary system that insures continued operation even if the primary system fails. The first definition for fail-safe (above) makes up the majority of the railroad concept of the term vital, and these two terms are sometimes used interchangeably, albeit somewhat loosely so, within the railroad signal world. However, in the U.S. railroad-signaling context, the definition of vital often combines the essence of fail-safe, as defined above, along with the additional implication that the circuits have specifically been designed in accordance with the requirements of the Federal Railroad Administrations (FRAs) rule book (U.S. Code of Federal Regulations, Title 49, Parts 233 through 236). The key to vitality for U.S. railroads lies in what the FRA refers to as the closed circuit principle. The closed circuit principle means that the relays and other logic devices used in safety-critical circuits must normally be energized. Furthermore, the de-energized state must also represent the most restrictive state of that portion of the signaling system. FRA Rules 234.203, and 236.5 Design of control circuits on closed circuit principle All control circuits the functioning of which affects safety of train operation shall be designed on the closed circuit principle, except circuits for roadway equipment of automatic train stop system.

5-3

Abnormal Operation of Railroad Equipment

For example, the simplest direct-current relay-based vital track circuit, used to detect the presence of a train within a particular section of track, is constructed using a track relay whose coil is normally energized by current transmitted via the rails from a battery at the opposite end of that section of track. In this case, normally refers to times when there are no axles of a train or railroad cars present (shunting the track) within the physical boundaries of the track circuit. Thus the coil of the relay is energized by the absence of a train within the block1.

Figure 5-1 Elementary Steady Energy DC Track Circuit

Because the signaling system is made to display its most restrictive indication (usually denoted by a red signal meaning STOP) to all other trains by the removal of energy from a device, and not by its application, a vital system will fail-safe when internal failures such as a broken wire, a short circuit, a cracked relay coil winding, or the complete failure of the power supply interrupt the normal flow of electric current. This sort of fail-safe or vital behavior will also occur in the case of a broken rail, or any other opening of the circuit that can de-energize the coil of the track relay. As the circuit is normally energized, this serves as a continuous form of self-testing. In addition to the open circuits that can de-energize the coil of a track relay, a track circuit will also respond to the presence of a short circuit. This short can be created either by the shunting of the track by a trains axles, or by any other form of short circuit, such as an insulation failure in the circuit wiring, or even metallic debris placed across the tracks anywhere within the track circuit. When the short is from a normal and expected source, such as the axles of a train, or a
1

When the track circuit is unoccupied, energy flows along both rails to the coil of the relay (connected at one end of the track circuit) from the battery connected to the track at the opposite end. When a train enters the track circuit, the axles of the train shunt the current being supplied to the relay coil from the battery, and this de-energizes the relay. This action causes the wayside signal system to display the necessary signal aspects (red, green, yellow, etc.) at various locations in both directions along the track, and this instructs other trains not to enter the already occupied sections of track. (See: Overview of Railroad Signal Circuits, Chapter 4 in this handbook.)

5-4

Abnormal Operation of Railroad Equipment

signal maintainers test shunt, this is usually referred to simply as a shunt2. When the shunt results from any unintended source, it is called a false shunt. The presence of a shunt, or a track circuit defect such as those mentioned above, should cause the system to fail-safe, because the circuit is designed on the closed circuit principle. This failsafe action normally takes the form of the system entering its most restrictive state. In the case of wayside signals, this usually means turning the signals at either end of the signal block to red, as well as changing the indications displayed by other more distant wayside signals. However, in the case of grade crossings, the behavior of the system is often more complex. As motion sensors and crossing predictors are designed to respond primarily to a moving shunt within the circuit, the mere presence of a shunt (false or otherwise) may not be sufficient cause to put the gates down and flash the lights. This depends primarily on the location of the shunt and the specific characteristics of the train detection equipment. But any single track circuit failure, which could prevent the railroad signal system from correctly detecting the presence of a train or axle(s) within the circuit must either cause the system to fail-safe, or be covered by a rule3 in order for the system to be considered vital. There are also several examples of regulations designed to prevent unsafe abnormal operation resulting from multiple failures (e.g. unintentional grounds in signaling circuits). Redundancy vs. Vitality Although redundancy is used in some parts of railroad signaling, vitality differs from simple redundancy or the use of supervisory safety circuits, in that each component or subsystem which is relied upon for proper operation of a vital system, is either vital, or is supervised by a vital circuit which monitors its operation. Even the gate mechanisms used at grade crossings are designed to be vital. If all power (including the backup batteries) should fail at a grade crossing, the warning lights on the crossing would not be lit, and the bells would not ring. But since the gate mechanisms themselves are vital, they are designed to immediately lower their arms in response to a complete power failure.

Unfortunately, the word shunt has more than one meaning on railroads worldwide, and personnel unfamiliar with the varied uses of this term should be careful to note the context within which it is used. Shunting is also the term used to describe the switching or movement of railroad cars, especially where they are being moved between a main line and a siding or spur track. This usage is particularly prevalent in the United Kingdom, but can also be occasionally heard on railroads in North America. Persons without a railroad background have been known to misinterpret statements by railroad personnel using this word, leading to unnecessary misunderstandings. Vital equipment may still have certain limited susceptibilities, which can result in unsafe abnormal operation. For example, Crossing Predictors are designed to provide consistent warning time for constant-speed trains. If a train approaches a crossing slowly, waits for the crossings warning devices to be activated, and then accelerates towards the crossing, a short warning time may be produced. For this reason, most railroads in North America adopt an operating rule that restricts the type of train movements that can be made while within 3,000 feet of a grade crossing equipped with such automatic crossing signals. There are also other rules from governmental agencies, which are designed to prevent unsafe abnormal operation of the grade crossing warning systems and wayside signaling systems resulting from causes that cannot simply be designed out of the electronics.

5-5

Abnormal Operation of Railroad Equipment

They will block highway traffic until the supply of power, and proper functioning of the train detection equipment, have been restored. In contrast, one of the classic forms of supposedly fail-safe systems used in other industries is to use redundant non-vital circuits. A supervisory circuit checks the operation of the primary system, and if it is determined to have failed in some way, control is transferred to a backup system. All of the subsystems and components used are of good quality, and high reliability, but none of them are vital. There is no self-testing done by either the primary or backup systems, as this responsibility is given to the supervisory circuit. But even the supervisory circuit will not be checking itself, and will often only check the function of the primary system.

Figure 5-2 Abnormal Operation: Classic vs. Vital Redundant Systems

Thus, a failure in the supervisory circuit could remain hidden until the primary system failed. This latent failure could then result in undesirable operation when the primary system failed, if the supervisory circuit took no action to transfer control from the primary system to the backup system. Vitality is not achieved simply through the use of high-reliability components, but through an architecture that relies on the continual or continuous verification of the health and functionality of all components essential to safety. Dual vs. Vital This distinction between redundancy and vitality sometimes gets muddled when the equipment being discussed actually consists of two complete and separate vital systems constructed in a single chassis. Both major U.S. manufacturers of Crossing Predictors and Motion Sensors offer a dual version of their equipment. Such designs not only include a complete spare unit in a single chassis, they also have a transfer or changeover interface which connects only the active unit to the inputs and outputs of the chassis. Thus, when a persistent right-side failure occurs within the primary unit, control is transferred to the secondary unit.

5-6

Abnormal Operation of Railroad Equipment

The net effect of this is to create a crossing warning system which, when an internal failure does occur, will be falsely activated (right-side failure mode) for a period of only a few minutes before control is transferred to the secondary unit. Without this secondary unit, the crossing would be falsely activated until signal maintenance personnel arrived to make repairs. Most dual systems like this are installed at locations where the potential costs of having a crossing down unnecessarily for more than a few minutes are very high. Crossings located far from the maintainers headquarters, busy intersections, and crossings frequented by impatient drivers likely to smash through the gates with a large truck, all fall into this category. These systems are also attractive from a maintenance perspective, as their built-in spares make the return of a failed module or sub-system to the factory for service much easier. Control of the crossing is manually switched to the good unit, while the failed module from the bad unit is removed for repair. This temporarily turns the dual system into a single system, but otherwise there is no change in performance. Each vital system incorporated into these dual systems still has the degree of internal redundancy and self-checking needed to be called vital in its own right. As always, vitality does not come from sheer redundancy, but rather from a normally energized, closed-loop, self-checking, fail-safe design. Abnormal Operation Now that we have explained some of the underlying terminology of railroad signaling, we are now ready to define and discuss Abnormal Operation. We will define abnormal operation, for railroad signaling systems, as a significant departure from the normal operational behavior of the signaling system, which results from one of the following causes: 1. A wrong-side failure mode 2. A right-side failure mode 3. spoofing of normal events.
Wrong-Side Abnormal Operation

Abnormal operation resulting from wrong-side failures is extremely rare. The primary reason for mentioning wrong-side failures here is to point out that in the case of severe railroad signal equipment damage resulting from a catastrophic event, multiple vital component failures may be created in the signaling equipment simultaneously. This most often results from lightning, surges, or power line faults that cause excessive voltage to be applied to the inputs or outputs of the signal equipment. Although there are several layers of protective devices installed in railroad signal equipment, the external surge protection devices used are primarily designed to provide protection from nearby lightning strikes, and may not be able to withstand the longer durations and larger energies associated with power line faults. 5-7

Abnormal Operation of Railroad Equipment

In the event that multiple failures are created within the signaling equipment in a simultaneous or near-simultaneous fashion, a wrong-side failure may occur. In many ways, this is no different than what would be expected to happen in the event of a fire, flood, or other environmental catastrophe that creates conditions that are outside of the equipments intended operating environment. Most likely, the equipment will fail-safe, due to inherent redundancies and the use of the normally-energized principle in the design of the equipment. But almost anything is possible after multiple simultaneous failures. This is because the equipment is only required to be singlefault tolerant. But in many cases, manufacturers build in multiple fault tolerance for some types of failures, just to be on the safe side. The term: wrong-side failure means that a deviation from normal signal system operation has occurred, the signal system has not detected it, and this has resulted in a potentially unsafe condition or mode of operation. As the purpose of most railroad signaling equipment is to provide an indication of what movements may safely be made, both by trains and the motoring/pedestrian public, the end result of this failure is usually some form of a false clear signal aspect, which gives the wrong indication to the train operator or public. The term false clear comes from the terminology used to describe wayside signal aspects. The proper name for a green signal aspect is clear. If a train has a clear signal, this means that they are looking at a signal that is usually green in color, and tells them that it is safe to proceed onto the track(s) ahead, at or below the speed limit listed in the railroad timetable. Unlike the old If a tree falls in a forest. . . conundrum, the necessary and sufficient conditions of a wrong-side failure or false clear are well-defined. Specifically, a false clear is a failure of the signal system only, and is still said to have occurred, regardless of whether or not there was anyone (or any train crew) around to see it. A wrong-side failure usually takes one of two forms, either: 1. The signal system has displayed a more permissive aspect than was safe or proper. For example, a green signal aspect was displayed, when the signals aspect should have been either yellow or red. 2. The warning devices at a grade crossing were not activated when they should have been. Examples of this include: short warning times, non-existent warning times, and crossing warning devices shutting off while the train is still occupying the crossing. Wrong-side failures are the quintessential nightmare of all railroad signalmen and signal equipment designers. These events, although rare, are cause for a thorough investigation into the root cause(s) of the failure, and (in the United States) must be reported to the FRA. Wrong-side failures can result from something as simple as the mis-adjustment or mis-wiring of the track circuit. This is why in the United States, the FRA requires testing of all vital circuits at the time of installation, periodic re-testing of all vital signal circuits, and requires that a vital circuit be re-tested any time that more than one wire has been disconnected at the same time 5-8

Abnormal Operation of Railroad Equipment

during the course of routine maintenance or repairs (called a disarrangement). By the enforcement of good policies, and the use of good engineering practices, the risk of wrongside failures can be minimized, although not completely eliminated. Most wrong side failures result from a much more mundane source: human error. Failure to completely and correctly test a signal system at the time of its installation, or the failure to maintain and repair it correctly, can result in wrong-side failures. Like ac power systems, railroad signal systems are designed to operate continuously. The time required for repairs to the equipment installed at grade crossings can easily exceed the average motorists patience. Thus, systems are routinely jumpered out using temporary wiring to keep the warning devices de-activated while repairs are being completed. These methods are used only after taking measures to provide for the safety of highway traffic, such as by obtaining track and time from the dispatcher, in order to temporarily halt the flow of trains. But the danger often arises after the work has been completed, and the maintenance personnel have left the area. Lives have been lost simply because someone left a jumper wire or wires connected to the equipment after they were done working on it. Even though the signal equipment itself was working properly, the temporary jumper(s) left behind prevented the warning devices at the grade crossing from working when the train approached the crossing, thus resulting in tragedy.
Right-Side Abnormal Operation

A right-side failure means that some significant deviation from normal signal system operation has occurred, but there has been no reduction in safety. This mode of operation may be due to a physical component failure in the signaling equipment, or a software bug, or some other cause. Regardless of the root cause, right-side failure modes are by far the most common cause of abnormal signal equipment operation. Most commonly, right-side failures result from external track-related problems such as low ballast resistance, a broken track wire, a broken rail, or some other readily identified physical failure. The technical support staff at one major railroad signal manufacturer has a saying: Ninety percent of everything is track., referring to the overwhelming percentage of trouble calls that they receive which have nothing to do with their equipment, and instead stem from a failure in the wiring, rails, or track ballast to which their equipment is connected. When modern signaling equipment enters a right-side failure mode, this means that the system has either detected an actual failure, or is otherwise unable to confirm its own proper functioning. The occurrence of a right-side failure means that the system has both detected the presence of a problem, and has taken appropriate action to ensure the continued safety of both train movements and the public. This action usually takes the form of placing the system in its most restrictive state, i.e. the wayside signals for the trains turn red, or the warning devices (gates, lights, bells, etc.) at a grade crossing are put into operation. The signal equipment will remain in this condition until the failure goes away, or possibly longer, due to the use of delay timers which prevent rapid back-and-forth transitions between the normal and failure modes of operation in the case of intermittent failures.

5-9

Abnormal Operation of Railroad Equipment

In order to be vital, railroad signal equipment must enter a right-side failure mode whenever the proper functioning of any vital part of the signal system cannot be verified. This means that the equipment is intentionally designed to produce red wayside signals and false crossing warning system activations in response to the presence of poor ballast resistance, the presence of a track or wiring defect, or to the presence of interfering electrical signals on the rails. As inconvenient as this behavior may be, there is presently no safe alternative. So while right-side failures can be considered part of a systems normal response to interference, we will, for purposes of this discussion, continue to refer to them as abnormal operation in order to signify the departure from normal signal equipment operation. Signal equipment manufacturers often refer to such right-side failure modes as errors, or error conditions. These error conditions do not always indicate the presence of an internal hardware or software failure within the equipment itself. Rather, they usually indicate the presence of unusual external conditions or problems. When discussing modern electronic track circuits, saying that an error has occurred means that a right-side failure has occurred (abnormal operation), and this does not mean that unsafe abnormal operation has occurred. Some signal equipment has a built-in memory or data recording capability, which allows it to record the occurrence of an error condition. This may include a time stamp as well as other information. This is not unlike the flight data recorder of a commercial airliner. Unfortunately, the recorded error will seldom consist of a clearly worded text message telling the railroad signal maintainer the exact cause of the abnormal operation. Instead, the message will usually indicate what subsystem was affected, and in what way. This information may or may not lead the maintainer directly to the cause of the problem. As there are many potential triggers for each error condition that can be detected by the signal equipment, the first task is to determine the exact cause of the problem (see chapters 9 and 10). AC interference is sometimes the cause of abnormal operation of railroad signal equipment, but there are many other more common causes. The presence of ac interference should never be assumed to be the primary suspect until there is either direct evidence of the presence of excessive levels of ac interference voltage present between the rails of an abnormally operating track circuit, or all other potential causes of the observed symptoms have been eliminated. Perhaps the most classic example of the abnormal operation of railroad signal equipment, where ac interference is implicated, is having the warning devices activated at a grade crossing when there is no train approaching. As the most susceptible receptors of ac interference, the Motion Sensors and Crossing Predictors installed at grade crossings are usually the first indication of an ac interference problem. Under conditions of moderate to severe interference, the presence of ac interference energy (usually 60 Hz and its harmonics) can fool the motion sensor or crossing predictor into believing that a vital failure has occurred. This happens when the noise level rises high enough to distort the measured characteristics of the electrical signals transmitted and received by the Motion Sensor or Crossing Predictor. In order to perform their intended function, and to make sure that they are doing so in a vital fashion, many different forms of signal processing and measurement are performed by modern 5-10

Abnormal Operation of Railroad Equipment

train detection equipment. These include the measurement of the amplitude, frequency, phase, and modulation of the electrical signals transmitted and received by the Motion Sensor or Crossing Predictor. Any noise that interferes with this process can make the signal equipment believe that it is either no longer transmitting the signal onto the tracks correctly, or that it is not receiving and processing the electrical signal properly. This will result in the equipment entering a right-side failure mode. In general, the most consistent characteristic of right-side failure mode abnormal operation is the presence of self-diagnosed error conditions resulting either from actual hardware failures, or from interference.
Abnormal Operation from Spoofing of Normal Events

The third possible cause of abnormal operation, results from the signal equipment misinterpreting ac interference as a legitimate electrical signal conveying specific information about the presence or motion of a train. This often occurs when the level of interference present on the railroad tracks is not high enough to induce an error condition (see above) resulting in a right-side failure mode, but is still high enough to interfere with normal operation of the equipment. Specifically, ac interference voltages on the rails (ac rail-to-rail potential) that varies with time in amplitude, frequency, phase, or a combination of these, can mimic or spoof the electrical signature of a train. The signal equipment then responds to this illusion as though it were an actual train, often putting the gates down or declaring the track circuit to be occupied when it really isnt. In these cases, the random nature of the interference plays a key role in a statistical game. Whenever the interference matches the electrical signature of a train, the illusion will be created. However, the illusion will usually not persist over time, and the most common result of this type of interference will be a grade crossing whose warning devices go into operation for short periods of time at random intervals. This apparently random crossing warning system activation is often a function of the amplitude of the interference, with the crossing being activated more frequently as the level of interference rises. In order for induced ac voltages to affect railroad signal equipment, there must often be a failed, damaged, or externally-influenced component to act as a catalyst for the ac interference energy. Shorted insulated joints commonly fill this role, but other signal system components can also exhibit this behavior. For an example of how this can occur, consider the following: The presence of high levels (a few volts, or a few tens of volts rms) of 60 Hz energy across the track (as measured from one rail of the railroad track to the other) can detune some narrowband termination shunts. These termination shunts are normally connected from one rail to the other at the physical limits of the track circuits used by motion sensors and crossing predictors. Consisting of a series-resonant inductor-capacitor network, these termination shunts effectively short the rails together at a particular frequency (usually in the range of 80-1000 Hz). But these networks are not perfectly frequency-selective, and thus may also pass significant amounts of interference current at other frequencies, if the rail-to-rail voltage is high enough.

5-11

Abnormal Operation of Railroad Equipment

If the amount of interference current passing through the termination shunt is sufficient to cause significant saturation of the magnetic core of the termination shunts inductor, the shunt will become de-tuned. This occurs because the inductance of a saturated inductor drops below its nominal value. This re-tunes the series-resonant LC network to resonate at a new, higher frequency. If the amount of detuning was sufficient to move the center frequency of the shunt so that it coincided with the frequency being used by another audio-frequency track circuit (A.F.O., motion sensor, crossing predictor, etc.), the frequency-shifted termination shunt could, in theory, shunt the track effectively at that frequency. This could potentially shunt enough current away from the receiver of an audio-frequency overlay track circuit to cause the A.F.O. to declare the track to be occupied by a train. And if this action proceeded at the proper rate with respect to time, it could also fool a Motion Sensor or Crossing Predictor into believing that there was a train approaching a grade crossing. This form of track circuit interaction under the influence of ac interference voltages on the track can be difficult to find, and requires a broad knowledge of the particular characteristics of each signal system component in the presence of ac interference voltages. The hallmarks of the spoofing type of abnormal operation are that grade crossings will activate their warning devices for no apparent reason at essentially random intervals, and track circuits may declare themselves occupied in a similar fashion. Moreover, all of this will occur without the presence of recorded error conditions, or other evidence of right-side failure modes (in signal equipment equipped with error-recording capability). Philosophy Although the underpinnings of railroad signal design philosophy are simple, their implications are manifold and complex. At their most basic, railroad signaling systems are designed to facilitate the safe movement of trains, and are vital and fault-tolerant to at least the single-fault level. Actually achieving this level of performance requires an enormous amount of effort and expense, and no small amount of time. There are many ways of meeting these requirements, but with little in the way of regulation or customer-based requirements to specify the details of equipment design, signal equipment manufacturers are largely left to their own devices to fulfill them. However, this does not relieve them of the need for vitality and the use of due diligence in the design of their equipment. Divergent opinions as to exactly what practices are required in order for equipment to be truly vital are held by many respected experts in this field. Some arguments about the specifics of signal equipment vitality ultimately become what one signal equipment engineer has termed philosophical, meaning that they are based on what an individual engineer personally believes is necessary to protect against foreseeable events. Other arguments about equipment vitality are more statistically based, especially with regard to the timing of multiple failures in electronic systems. Even though fault-tolerance is only required to the single-fault level, some vital designs which do not self-test in a truly continuous fashion have been analyzed with the supposition that an initial vital failure could be closely followed in time by a second failure, which could theoretically prevent detection of the first failure. Therefore the operational behavior of some signal equipment is discussed in terms of a vanishingly small mean time before wrong-side failure. 5-12

Abnormal Operation of Railroad Equipment

That any wrong-side failure could ever occur in truly vital signal equipment is an oxymoron, but the introduction of computer technology into railroad signaling systems has, in effect, forced the expansion of the philosophy of railroad signal system vitality to include sampling-based systems which process information in a discrete, non-continuous fashion. In general, these systems seek to perform a self-test frequently enough so as to detect any vital failure within the normal amount of time required to detect a train. Methods of Achieving Vitality Vitality in railroad signal equipment has been achieved in many different ways. The most basic technique used is the normally energized principle discussed above. Another common technique for electronic systems is that of continuous, or intermittent-but-frequent, self-checking. Modern electronic track circuits may also achieve vitality through multiple independent signalprocessing channels, whose outputs are continually compared with each other by an independent supervisory system. The same inputs are applied to both signal-processing channels. If there is sufficient disagreement between the outputs of the two channels, then the supervisory circuit declares some part of the system to have failed, and a right-side failure mode is entered. Thus, a failure that affects either signal-processing channel in a vital fashion will cause the system to fail-safe. The objective is to guarantee that the introduction of any single failure into the system, which could affect the safety of the system, will cause the system to fail-safe. This is called singlefault tolerance. Single-fault tolerance means that the failure of any single component, process, system, or sub-system, which can adversely affect the safety of the system, will cause the system to enter a right-side failure mode. Vital railroad signal systems having single-fault tolerance must also have no latent failures that could affect the safety of the system. That is, a vital failure must always make its presence visible or known. Normally this is done by entering a right-side failure mode that activates the grade crossings warning devices, or sets wayside signals to red. This immediate self-revealing behavior is necessary, as a subsequent additional failure could result in unsafe operation (wrongside failure mode). Even though the system is single-fault tolerant, it is not completely foolproof, and may still behave incorrectly in the case of multiple simultaneous or nearly-simultaneous failures. The occurrence of multiple simultaneous failures is clearly beyond the realm of single-fault tolerance, and may cause a wrong-side failure to occur. However, the safety of modern railroad signaling systems is actually much better than mere single-fault tolerance would suggest. This occurs for two reasons: 1. The simultaneous occurrence of two failures, affecting both the safety of the signal system and the ability of the system to detect this particular failure, is extremely unlikely. 2. Because the probability of item 1 (above) is not actually zero, signal equipment manufacturers have worked diligently to cover as many scenarios involving multiple simultaneous failures as possible, even though this is not explicitly required by law or customer specifications. It is simply good engineering practice, and good risk management from a business perspective. 5-13

Abnormal Operation of Railroad Equipment

The number of ways in which a system can be made vital is limited only by the number of ways in which the system can fail. Each component, sub-system, or system has its own unique susceptibilities. For each potential hazard or wrong-side failure mode (see below) which could occur, mitigation measures must be devised which either prevent the failure from happening, or continually test (in a vital fashion) the proper functioning of that component, sub-system, or system, and initiate a right-side failure mode whenever self-testing reveals a problem. Vitality and Creative Equipment Maintenance One important sidebar to this discussion is the topic of signal equipment maintenance. This issue must be raised in order to warn against some of the more creative solutions that have been attempted in an effort to improve railroad signal equipment operation. Vital signal equipment designs can be very intricate. As with many fields of engineering, the devil is in the details. Creating a vital design consists of making sure that the effects of every potential failure or hazard are well understood, and properly mitigated against. Its not unlike making sure that youve thoroughly plugged every hole in a large and porous dam. This method of design can be susceptible to the effects of seemingly insignificant modifications. One small leak is all it takes. For example, some vital designs rely on the use of specially designed and manufactured components, simply because they have limited and predictable failure modes (vital relays, for example). Even though a conventional dc track circuit could theoretically be built using a garden-variety relay, purchased from the local electronics store at a significant cost saving. However, the circuit constructed would not be vital, owing to the inexpensive relays use of allmetal armature contacts, which can become welded together if they are made to carry excessive current. Vital railroad signal equipment is most often returned to the manufacturer for repairs, but may also be serviced by factory-trained technicians in the shops of larger railroads, using manufacturer-approved parts. This tendency for conservative repair policies is understandable, given the safety-critical nature of this equipment. But this conservative approach applies to the introduction of new equipment and installation techniques as well. If you start a conversation about ac induction and interference on the railroads with an electrical engineer who is unfamiliar with railroad signaling systems, the first quick fix that they will usually suggest is to simply ground both rails. Oddly enough, this has been tried by wellmeaning industrial plant electricians who were installing grade crossing warning equipment on a spur track within the facility. Their familiarity with standard wiring practices for commercial power overcame the fact that the correctly-drawn signal drawings did not show any ground rods connected directly to the rails. They figured that any exposed conductor should be carefully grounded for safety, and so they did. Fortunately, in this particular instance, the location and relatively high resistance of the ground rods was such that they did not create a false shunt across the tracks within the approaches to the crossing. Had things been otherwise, this could easily have resulted in the crossing warning system producing little or no warning time. Furthermore, the behavior of the system could have been radically changed by the amount of moisture in the soil. This example points out the need 5-14

Abnormal Operation of Railroad Equipment

for caution in applying non-standard solutions to ac interference problems. Even something as apparently harmless as a ground rod can have very serious unintended consequences in railroad signaling. Creative maintenance is often a shortcut to abnormal operation of all kinds. Rules, Vitality, and Wrong-Side Failures Part of the vitality of the railroad signaling systems in North America resides not in the equipment itself, but in the practices surrounding the equipment. Policies must be in place to cover circumstances that cannot be adequately addressed by the signal equipment alone. It is for this very reason that some railroads require that all jumper wires used to jumper out the equipment at a grade crossing be at least six feet long, and constructed of wire with brightlycolored insulation. However, there are other circumstances that are much more common, particularly in some parts of North America. For example, if the external power supply (usually from the ac grid) fails at a wayside signal location, then eventually the batteries will become exhausted. This occurs most frequently after a heavy winter storm, at the most snow-bound and remote locations. Even with a large battery bank, an ac outage of several days can still drain the batteries, especially at remotely operated track switch locations. Such a complete power failure can ultimately result in one or more dark (un-illuminated) signal heads. This failure cannot be easily remedied with additional technology, and therefore is covered by operating rules. The presence of a dark signal head is cause for the trains engineer to stop the train, report the problem to the dispatcher, and not to proceed until obtaining permission to do so from the dispatcher. This policy also ensures that trains do not pass a signal that should be displaying a red aspect, but have a burned-out bulb in them. Of course, this also relies on the trains engineer being very familiar with the territory when operating at night. If such a power failure were to happen at a grade crossing, it would eventually result in crossing gate arms that are horizontal, but have no flashing lights on them. Such situations must be addressed by the use of an FRA rule, railroad operating rule, or motor vehicle code. The safety of the public is still protected, because a lowered gate arm, even without any lights, is still an indication that the grade crossing should be considered temporarily impassable. There are also a few exceptions to the single-fault tolerant nature of vital railroad signal equipment which have wrong-side failure implications. An example of this is the presence of a false shunt across the tracks within the approach(es) of a Motion Sensor or Crossing Predictor. Such a false shunt could blind a Motion Sensor or Crossing Predictor to the presence of an approaching train. In practice, this sometimes takes the form of metal debris (shopping carts, pipes, scrap metal, etc.) placed across the tracks in the path of a train by vandals. The debris is not always knocked clear of the rails by the train, and sometimes remains wedged between the rails. In such cases it is the rules regarding red wayside signals which provide for continued safe train operations, as the false shunt created would also keep the wayside signals in their most restrictive state (red or stop) prior to a trains arrival. The rules for most North American railroads severely restrict the manner in which a train can be driven past a red signal, and absolutely prohibit driving a train past some red wayside signals (absolute signals) without first obtaining permission from the dispatcher. Thus, any train approaching such a false shunt would, as per the 5-15

Abnormal Operation of Railroad Equipment

rules, be running at restricted speed, and would be approaching each railroad crossing prepared to stop before reaching it. There are also other rules from governmental agencies, which are designed to prevent unsafe abnormal operation of the grade crossing warning systems and wayside signaling systems resulting from other causes that cannot simply be designed out of the electronics. All of these regulations are designed to avoid the abnormal operation of railroad signaling equipment. Railroad Signal Equipment Standards vs. Other Industries The intentional design of railroad signaling equipment to exhibit a right-side failure mode in response to anything that could adversely affect the safety of the system has resulted in surprisingly few customer requirements limiting this behavior. Ideally, a piece of signal equipment would only enter a right-side failure mode when absolutely necessary, and one might expect the development of customer requirements which specify a maximum false-alarm rate for a given environment. However, unlike some other fields of electronic signal processing, there are no widely accepted standards within the U.S. railroad signal industry defining a maximum allowable equipment failure rate, or a maximum false-alarm rate for a given interference environment. Equipment acceptability has instead been determined by the railroads on a rather informal basis. When new equipment is introduced by a vendor, site trials are often set up at a limited number of railroad signal installations with the new equipment operating in shadow mode, if possible. In this form of test, the existing equipment and its proposed replacement are placed in operation side-by-side on the same track, if possible. The existing equipment remains in full operation, with the new equipment only connected to the required track wires, power sources, etc. needed for operation. But the new equipment is not allowed to control anything. The outputs of the new and the existing equipment are both connected to a data or event recording system, and the equipment is left to operate for a period of time. At the end of the test period, the contents of the recording system are downloaded and analyzed. The performance of the new equipment is compared with that of the existing equipment, and if the new equipment performs at least as well as the existing equipment, it may be approved for use on that railroad. In these trials, other factors such as purchase price and long-term support costs often play a larger role in the decision than performance, provided that the new equipments performance is at least as good as the old. A thorough statistical analysis of the performance of the new equipment is rarely done, and the results of these site trials are hardly ever translated directly into formal numerical performance requirements for future equipment purchases. Instead, equipment which proves to be unreliable in service quickly gathers a bad reputation, and railroad feedback is provided to the signal equipment vendor via the vendors thoroughly abused sales and technical support staff. Thus, detailed standards for railroad signal equipment performance, including its abnormal operation rate, have seldom, if ever, been developed in North America.

5-16

6
DAMAGE TO RAILROAD EQUIPMENT
This chapter describes damage to railroad equipment that can result from electromagnetic interference. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA.

6-1

Damage to Railroad Equipment

Introduction
Damage to railroad signal equipment is one possible outcome of ac interference. This chapter examines the mechanisms of equipment damage from ac interference, and design practices that can affect equipment survival.

Quick-Start Version
When railroad signal equipment damage is caused by ac interference, it is usually fault or switching activity on the power system that causes the problem. Steady-state ac interference doesnt often cause damage because railroad signal equipment is designed to withstand ac voltage levels well above those considered hazardous to personnel (50 volts rms steady-state). Since steady-state interference levels are usually maintained below this level for personnel safety, steady-state interference rarely causes damage. (Notable exceptions to this are wideband shunts and narrowband couplers that can burn out if they pass too much steady-state ac interference current.) But track circuit arresters are designed to protect against lightning, not power line faults. Surge protective devices (SPDs or arresters) used on railroad equipment are designed to withstand only lightning. SPDs used on track circuits are generally designed to fail open (not shorted). And they 2 fail when the energy flowing through them exceeds their i t capacity (i=current in amperes, t=time in seconds). While the current through an arrester will often be an order of magnitude higher for lightning, the duration of the current can be three to five orders of magnitude greater for a power line fault. Thus the energy from induction from a power line fault can be much greater than from a lightning strike. Track SPDs often cannot withstand power line fault induction because they were never designed for that purpose. The result is that the arrester is often destroyed, leaving the rest of the railroad signal equipment vulnerable.

Detailed Version
Although a small amount of ac interference energy, related to the power-line frequency, is always present on railroad rails due to induction, it is usually inconsequential, and has no visible effect on the operation of the railroad signaling equipment. Somewhere above this normal level of ac interference lies a range of energies which can cause operational problems for the various types of signaling equipment in use, but will not cause any permanent damage. The exact voltage and/or current thresholds at which this occurs will vary widely, depending primarily on the type of equipment installed. However, these thresholds are also strongly dependent on other factors, including the operating frequency used by the equipment, the operating mode or modes selected, and the particular electrical characteristics of the track and associated components installed at each location. In general, as the level of interference rises, so does the rate of operational problems with the signaling equipment; in some cases this trend may only be visible through the use of statistics. As problems from ac interference are often random or pseudo-random in nature, the 6-2

Damage to Railroad Equipment

specific time(s) at which a signal system experiences trouble (assuming a constant interference level) will often be random or pseudo-random as well. But the frequency and duty cycle of visible symptoms will generally correspond to the level of interference - once the level of interference has risen above the threshold required to induce the first operational problem. Therefore, probabilistic predictions are often the only ones which can be accurately made about the performance of railroad signal equipment in a given interference environment. In the most extreme cases, the magnitude of the ac interference is so great that not only is the railroad signaling equipment prevented from operating normally - it is permanently damaged. This section discusses some aspects of this type of damage. Standards The field of railway signaling has very few standards for immunity to ac interference, either radiated or conducted. Railroad signal equipment manufacturers comply with the guidelines provided in the AREMA C&S Manual, part 11.5.1, Section D Environmental Limits, paragraph 6, shown below: 6. Electromagnetic Interference (EMI) Equipment shall operate normally without requiring adjustment when in service with all its covers on and installed per manufacturers recommendations at specified EMI field strength levels. Specified field strength is for either single impulse or continuous wave EMI. Processor-based equipment shall conform to requirements set forth in Federal Communications Commission (FCC) Rules, part 15, for spurious RF emissions with all its covers on and installed per manufacturers recommendations. This standard covers the ability of the system to withstand radiated interference received from the outside, as well as the limits imposed on the amount of interference radiated from each unit. Furthermore, it is one of the very few standards viewed by many in the railroad signal industry as having any relevance to the railroad signal equipment market in North America. The specified EMI field strength levels which are referred to by the AREMA standard above are as shown below in Table 6-1. The incoming ac interference levels specified in these standards normally are much less than those required to inflict permanent damage to the equipment. Furthermore, these standards cover only the radiated electromagnetic interference that the signaling equipment must withstand without operational difficulties, and they say nothing about the conducted interference that the signaling equipment must be able to handle. Although it is the conversion of an electromagnetic field (usually from ac power system related sources) into conducted interference (on the railroad rails) that usually results in operational problems, the level of conducted interference that such a field could normally produce is still normally far below the level needed to cause permanent damage to the equipment. The only significant exception to this generalization is, of course, that of a severely unbalanced track circuit (more on this below). 6-3

Damage to Railroad Equipment

Also, as the lower frequency limit of the EMI given in the AREMA specification is only 50 KHz (see Table 6-1), this standard does not cover the portion of the electromagnetic spectrum that is most likely to cause operational interference or permanent equipment damage.
Table 6-1 Damage Table, Taken from AREMA C&S Manual Part 11.5.1
Class A (Roadbed) Class B (Wayside Outdoors) Class C (Wayside Signal Enclosures) -40C (- 40F) +70C (+160F) -55C (- 67F) +85C (+185F) 0 95 0 95 0.07" p-p 1.5 g p 10 g p 10 g p 150 250 350 Class D (Wayside Control Room) -25C (- 13F) +70C (+160F) -55C (- 67F) +85C (+185F) 0 95 0 95 0.05" p-p 1.0 g p 10 g p 10 g p 150 250 350 Class E (Computer Room)

Parameter

Condition

Temperature

Relative Humidity (%) Noncondensing

Vibration Mechanical Shock (11ms) EMI (V/m)

Dielectric Strength Abrasive Environment

Operating Minimum Maximum Storage Minimum Maximum Operating Minimum Maximum Storage Minimum Maximum 5-20 Hz 20-200 Hz Shipping Operating 50 kHz-88 MHz 88-216 MHz 216-1000 MHz Volts RMS Salt, rain, sand, hail, dust, contaminants

-40C (- 40F) +70C (+160F) -55C (- 67F) +85C (+185F) 0 95 0 95 0.2" p-p 4.2 g p 10 g p 10 g p 150 250 350

-40C (- 40F) +70C (+160F) -55C (- 67F) +85C (+185F) 0 95 0 95 0.1" p-p 2.0 g p 10 g p 10 g p 150 250 350

+20C (+68F) +25C (+77F) -15C (+ 5F) +50C (+122F) 40 45 5 95 0.05" p-p 1.0 g p 10 g p ---150 250 350

3000

3000

3000* 2000** No

3000 2000** No

600

Yes

Yes

No

Note: * 3,000 Vrms for Electromechanical Equipment ** 2,000 Vrms for Electronic Equipment

6-4

Damage to Railroad Equipment

Thus, it is entirely possible for a piece of signaling equipment to meet all of the U.S. railroad signal industry standards currently in use, and still not work correctly in its intended interference environment. However, since signal equipment that doesnt work also doesnt sell very well, most signal equipment manufacturers strive diligently to make their equipment as robust, reliable, and useful as possible. To date, the railroad industry has generally relied on signal equipment manufacturers to use their best engineering judgment with regard to ac interference immunity, and has relied on market forces as its primary enforcement tool. There have been some initial and tentative attempts by industry working groups to grapple with this lack of standards for conducted interference, but at present there is little in the way of published and accepted EMI standards for either the ac power industry or the railroad industry to use in resolving conflicts between railroad signal systems and ac power lines. Railroad Signal Equipment Design Practices Although 60-Hz energy is below the frequencies used by most railroad signaling systems, it still falls within the roughly 0 to 1000 Hz audio-frequency band used by many of the most susceptible railroad signal systems (e.g. Crossing Predictors, Motion Sensors). And it is quite close in frequency to the 40-Hz or 60-Hz Cab-Signaling systems used on some railroads. This is why most of the standard railroad signal equipment operating frequencies (e.g. 86 Hz, 114 Hz, 156 Hz, etc.) have been chosen specifically to avoid 60 Hz and its harmonics. Higher train speeds, and efforts to reduce the number of signal blocks, have dictated the use of increasingly longer track circuits. This in turn has favored the use of lower signal equipment operating frequencies. So for the majority of Crossing Predictors and Motion Sensors, as well as many AFO circuits, 60 Hz energy lies well within one decade, in frequency, of the signal equipment operating frequency in use. With such a strong source of potential interference being located so close (in frequency) to the operating frequencies of many widely-used types of railroad signal equipment, some potential for interference and/or damage is almost unavoidable. Because of this, the coupling networks used by the signal equipment manufacturers are often designed to be of a highly frequencyselective nature. Tuned inductor-capacitor networks are often used as a track coupler for signal equipment designed to operate at a single frequency. This may have been a part of the equipment design simply to provide proper impedance matching, but this bandwidth-limited transfer of energy between the equipment and the railroad track can also help to minimize the potential for equipment damage. As long as the equipment operating frequency is far enough away from the significant spectral components of potentially damaging interference (usually 60 Hz and/or its harmonics), the possibility of abnormal operation or damage is generally lessened. But modern Crossing Predictors and Motion Sensors, which can be user-programmed for operation at a variety of frequencies, must have a track-coupling network with much greater bandwidth than equipment tuned for a single frequency. These systems can generally be programmed to operate at frequencies from a few tens of Hz - up to nearly 1 KHz. This broader bandwidth can make them more susceptible to 60-Hz energy than single-frequency signal equipment. 6-5

Damage to Railroad Equipment

Sometimes special measures such as notch-filtering or other bandwidth-limiting schemes are employed to improve the operation of signal equipment in high ac noise environments. These can improve the signal systems operational performance and reliability under routine levels of ac interference; but these measures are not actually designed to reduce the risk of equipment damage during exposures to potentially-damaging levels of ac interference. Any benefit that they may provide in this regard is purely coincidental. Unfortunately, the limited bandwidth of railroad tracks as an electrical transmission line has necessitated the use of dc or low audio frequencies in railroad signaling and train detection, and there will always be some degree of susceptibility to 60 Hz energy and its harmonics for these systems, including the potential for damage. Despite all of this apparently inherent susceptibility, railroad signal equipment is really quite reliable, both in terms of its operational performance and its resistance to damage from ac interference. This can be attributed to three primary factors, which include: 1. a long-standing awareness of environmental factors such as lightning, 2. the generally well-balanced electrical characteristics of railroad tracks, and 3. the way in which electrical signals are applied to the rails. Much of standard railroad signaling practice in North America is designed to prevent, insofar as practical, permanent damage to signal equipment as a result of lightning strikes. However, damage arising from excessive levels of 60-Hz energy has been largely ignored, as the levels required to cause damage are rarely achieved on the tracks. Energy that is inductively coupled onto railroad tracks will generally affect both rails equally, provided that there is no gross electrical unbalance between the rails. This results in nearly equal rail-to-ground potentials on each rail at any given point along the track. Thus, the rail-to-rail potential of 60-Hz (or its harmonics) seen by the signaling equipment at any given point within a section of track is generally quite small. So even though the ac rail-to-ground potential of the each rail may be greatly elevated, this seldom bothers railroad signal equipment. This is because the overwhelming majority of railroad track circuits in North America impress their signaling voltages and/or currents across the tracks. That is, the signal is applied differentially to the two railroad rails of each track circuit, with no intentional referencing of the voltage or current to any common ground or return path. DC track circuits have batteries (and 110/220-Volt battery charger outputs) that are quite intentionally floated with no reference to remote earth ground. In part, this helps to prevent the occurrence of unsafe or wrong-side signal system failures. (A wrong-side failure is an unsafe condition where the signal system fails to respond properly to the presence of a train.) Audio frequency track circuits are also floated, for many of the same reasons. The coupling network used to connect such audio-frequency signaling equipment to the rails of a track usually consists of a capacitor and transformer, as shown below. 6-6

Damage to Railroad Equipment

Figure 6-1 Track Circuit Coupling Network

Therefore, until an insulation failure occurs between the windings of the transformer, or at some other point in the signaling equipment, any common-mode elevated rail-to-ground potential that has been induced is essentially invisible to the signal system. As specified in the AREMA standard 11.5.1 (above), the required breakdown voltage for electronic signal equipment is 2000 Volts R.M.S. Obviously, this is a voltage level seldom achieved on the rails in any steady-state fashion (excluding certain electrified railroad problems). But this specification is useful in that it does define a common-mode voltage boundary beyond which a signal equipment failure may occur, especially during transients, surges, or other non-continuous conditions which are capable of creating this condition. Exposures Obviously, the electromagnetic phenomena that can damage railroad signal equipment fall into one of the following two categories: 1. anything which raises the rail-to-ground potential significantly above the 2000 Volt insulation breakdown requirement of AREMA 11.5.1 (Note: this should also be enough to trigger the primary surge protection devices, a.k.a. lightning arresters.) 2. anything which raises the rail-to-rail potential enough to produce power levels within components used in the signaling equipment which exceeds their power ratings Category #1 (above) is most commonly associated with non-continuous conditions such as lightning strikes, power surges, transients, and the like. This may result in: 1. Dielectric breakdown of the connectors, printed circuit boards, or transformer insulation: this usually results in the energy finding a path to the grounded chassis of the signal equipment, and may cause multiple component failures along the way. 2. Dielectric breakdown of capacitors: coupling capacitors and filter capacitors, which are located very close (electrically) to the inputs and outputs of the system, are particularly susceptible. 6-7

Damage to Railroad Equipment

3. Failure of secondary surge protection devices: These surge-protection devices embedded in the various signaling equipment appliances often trip (as intended) during brief over-voltage or over-current conditions, and may become sacrificial1. 4. A Cascade of Failures: A momentary event within the ac power system can initiate a cascade failure. In such a case, the railroad signal systems surge protection devices, when activated simultaneously, serve to connect sequential signal blocks together electrically, resulting in a larger than normal induction problem due to the longer section of electrically continuous track created. Damage, as a result of exposures in this first category, is often characterized by physical changes in the signal equipment that are readily visible to the naked eye (components discolored, burned open, vaporized, etc.). But the type of damage which results from the exposures in category #2 is often much more subtle. Individual components can be damaged by a long-duration exposure to a less severe overload, and may fail in ways that are not readily visible. Among the failures that could occur within this second category are the following: 1. Dielectric breakdown of the coupling capacitor on the track side of the track transformer (in a transmitter or receiver). 2. Excessive current in the coupling capacitor or in the track winding of the track transformer (in a transmitter or receiver). 3. Excessive voltage applied to the terminals of the signal transmitter or receiver that triggers the secondary surge protection devices incorporated into the system in a catastrophic fashion. 4. Excessive power dissipation in a resistor connected across the track for the purpose of current or impedance limiting (found in ordinary track circuits, and some AFO track circuit transmitters and receivers). 5. Failure of a coupling device such as a tunable joint coupler or wideband coupler (capacitor) due to excessive voltage and/or current available across an insulated joint, where the coupling device is connected from one rail segment to the next around an insulated joint in a rail. (The usual application.) 6. Failure of a termination shunt (tuned or wideband) connected across the tracks due to excessive current. (For tuned shunts, this can occur when the frequency of a tuned shunt falls close to 60 Hz, or one of its harmonics.)

SPDs are intended to protect the down stream equipment from surges by diverting the energy to ground. They are expected to remain functional after the event has ended, and to provide subsequent protection from future surges. However, every component has a limit to the energy it can withstand. If the SPD is subjected to more energy than it can tolerate, it will be destroyed as it protects the equipment. This is expected, and is inherent in the sacrificial nature of the device.

6-8

Damage to Railroad Equipment

Vital systems are designed to be able to detect the failure of any single component that can affect the safety of the system, and are generally designed to avoid wrong-side failures even in the event of simultaneous damage to multiple components. But this does not guarantee that railroad signal equipment will avoid wrong-side failures in all cases of simultaneous multiple-component failures caused by excessive voltage or current from transients, surges, and the like. Although we have identified some of the most common types of specific failures, we have not attempted to attach any specific levels to them, as they are both manufacturer and component specific. But there are specific instances where a dc or ac (rms) potential of as little as 12 Volts from rail-to-rail can cause permanent damage. What is generally consistent across the industry, however, is a design philosophy that states: Destruction before wrong-side failure equals success. This principle is embodied in almost every signal equipment design. Obviously, the level at which destruction occurs varies widely, but manufacturers are always increasing the amount of abuse that their equipment can withstand in response to customer demands. Any signal equipment that acquires a reputation (deservingly or otherwise) for being easily damaged in the railroad environment quickly falls in sales. Mitigation In an attempt to minimize equipment damage, there are many specific designs, appliances, and wiring techniques used in railroad signaling. Various forms of lightning arresters are used as non-linear resistance devices (primarily spark-gap based) which are connected from rail-toground or from rail-to-rail to clamp excessive voltages to a level that the rest of the signal equipment can withstand. Secondary surge-protection devices are also incorporated into the design of the electronics. These primarily consist of non-linear semiconductor-based devices that work in much the same way. Other devices such as so-called Faraday cages are increasingly being explored to improve the isolation of signal equipment from the outside world. Manufacturers of signal equipment shelters are now offering several different implementations of a shielded arrester panel enclosure in an attempt to reduce the currents induced in clean wiring inside the equipment shelter by the dirty wiring leading in to the shelter from the track outside. However, large-scale studies of the effectiveness of these designs have yet to be conducted. To date, only preliminary studies such as that described in Lightning and Surge Protection Study RS-99-006 by R. Reiff et. al. at the Transportation Technology Center, Inc. of Pueblo, Colorado have been performed. Unfortunately, neither of the above techniques is effective against ac interference voltages that are present on the rails in a differential fashion (rail-to-rail potential) at levels too low to trigger the surge protection devices connected across the tracks, but high enough to cause damage over time. Therefore, other approaches must be considered. There are two basic engineering approaches to dealing with such ac interference voltages. The first is to increase the impedance of the path that the ac interference energy is following into the railroad signal equipment. This often takes the form of a series-resonant Inductor-Capacitor combination tuned to the operating frequency of the signal equipment. With a correctly designed coupler, very little impedance is added to the circuit at the operating frequency, but a great deal of impedance is added at 60 Hz and its harmonics. However, in any vital or fail-safe signal system, great care must be taken to assure that no potential failure mode of this coupler (open, 6-9

Damage to Railroad Equipment

shorted, out-of-tune, etc.) can cause a wrong-side failure. This restriction often prevents the use of this approach in certain types of signaling equipment. The second approach, often the first one employed by the railroads, is to try and shunt out the interference. This usually takes the form of series-resonant Inductor-Capacitor combinations, tuned to the offending frequency, and placed across the rails (in the same fashion as termination shunts), or from each rail to ground. Here the intent is to provide an alternate path for the ac interference, and to reduce the rail-to-rail ac interference potential seen by the signal equipment. However, these actually work by artificially restoring the rail-to-rail voltage balance on the track at a particular frequency. However, as the basic principles of ac induction and the role of track circuit unbalances in the production of rail-to-rail (differential) voltages are generally poorly understood by the railroad signal community, the use of 60-Hz Shunts to address high levels of ac interference on the rails usually relies more on luck than on science. It is often found that any level of ac interference high enough to cause damage will also be high enough to exceed the rated current carrying capacity of the 60-Hz shunt, which greatly minimizes the usefulness of this technique. This is not to say that these devices cannot be of benefit in specific applications, but these devices were designed and intended for use where the levels of ac interference were only high enough to cause operational problems with the equipment, and not high enough to cause permanent damage to the signal equipment (see above). Virtually the only effective forms of mitigation for ac interference rail-to-rail voltages high enough to damage railroad signal equipment is to attack them at their source, or modify their path to the affected equipment. Often, the most effective approach is to identify and correct the track circuit unbalance that is causing a difference between the voltages induced in each rail of the track. A reduction in the strength of the electromagnetic field, as measured at the rails, is also effective, but usually much more costly and more difficult to achieve.

6-10

7
PERSONNEL SAFETY CONSIDERATIONS
This chapter discusses in depth electrical hazards to personnel that could occur on a railroad as a result of power lines. This includes induced and conducted energy, as well as electric and magnetic field exposure and safe working clearances. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE. Mr. Frazier has over 30 years of experience in the management and conduct of applied research programs and the application of the research results to the solution of system-specific problems. His major fields of experience include electromagnetic effects, electromagnetic compatibility, cathodic-protection systems, and detection systems. He is the author of the EPRI computer program CORRIDOR that has been extensively used by the engineering community to predict power-line interference on co-located structures. He is nationally known for his work on the electromagnetic compatibility of electric power, pipeline, and railroad systems. As President of Corr Comp Co., Mr. Frazier provides services to the power/pipeline/railroad industries in the pursuit of compatible common utility corridors.

7-1

Personnel Safety Considerations

Introduction
Everything about operating a railroad or a power company is focused on safety. In a railroad environment, safety can mean many things. Safe train operations eliminate the accidents in which trains hit things, leave the tracks, or become involve in any number of other unacceptable events. Railroads generally do a very good job of this. With regard to railroad signals, ac interference can cause mis-operation of signals. An example of this would be crossing gates coming down when no train is coming. This would compromise safety. But there is a second aspect of safety involving ac interference. That aspect of safety, sometimes called personnel safety, involves the direct effects the electrical energy can have on a person. If the ac interference levels are high enough, a shock hazard may exist. This fundamental aspect of safety is just as important as operational safety, and hazardous electrical levels are just as unacceptable. Additionally, a third aspect of safety is exposure to high-level electric and magnetic fields. This will also be addressed along with various applicable standards for field exposure.

Quick-Start Version
It is not possible to list levels that insure safety in all situations and are economically feasible. For this reason it will be necessary for the reader to also study the detailed version of this chapter. Shock Hazard Electric Induction Where voltage is induced in railroad facilities by electric field induction, the steady-state short circuit current to ground should not exceed 5 mA ac rms1. This usually applies to high impedance communication circuits, large trucks and vehicles. It does not apply to cases of magnetic induction or conduction. Care must be taken when fueling machinery under high voltage lines. The power company should be consulted for specific recommendations. As a general rule, if fuel is to be transferred under high voltage power lines, the fuel container should be electrically bonded to the equipment being fueled prior to and during fueling. Any fumes should be allowed to dissipate before removing the bond. Magnetic Induction A minimum criteria for steady-state voltage induced on railroad facilities by magnetic induction would be to limit voltage to a maximum of 50 V ac rms point-to-point (within reach) under worst

National Electric Safety Code (NESC) among other standards.

7-2

Personnel Safety Considerations

case conditions2. However, this condition alone is not sufficient to insure safety in all situations. There will be circumstances where stricter criteria are appropriate. A thorough review of the detailed version of this chapter is highly recommended (as well as study of referenced sources). Fault Induced Voltages Computer modeling can be used to predict fault current magnetic induction into railroad systems. The same computer models can be employed to evaluate various mitigation options. But the target voltage values for safety under fault current induction need to be predetermined according to the site-specific situation. The 430 V rms and 650 V rms levels used by CCITT, the Canadian Standards Association, and the AAR/IEE Bluebook can be used provided the levels can be reasonably met with practical mitigation, and provided there are no unusual circumstances requiring lower levels. In any case, it is reasonable to evaluate the situations using the IEC 479-1 method or the IEEE Std-80 method (or both) to insure adequate safety and to insure mitigation is not unnecessarily expensive. The worst-case fault is modeled to determine the voltage. What constitutes the worst-case includes: Modeling various fault locations to find the one which produces the maximum voltage Fault the closest phase conductor to the track under elevated temperature final sag conditions Track circuit lightning arrester clamping ability may or may not be included in the worst case conditions by mutual agreement between the railroad and the power company.

Electric and Magnetic Field Exposure This topic does not lend itself to a brief summary. See the detailed section below entitled Electric and Magnetic Field Standards and Guidelines.

If point-to-point voltage is limited to 50 V, and an insulated joint has 50 V across it, then the rails would each be limited to 25 V to ground (assuming equal block lengths and uniform excitation).

7-3

Personnel Safety Considerations

Detailed Version
Shock Hazard This section of the Handbook will explore the question of what constitutes a shock hazard and what Standards exist that we can use as guides. Unfortunately, the debate continues regarding this issue. As a result you will see more than one possible conclusion that can be drawn from this examination of the topic. While the authors have their own opinions, all relevant Standards and opinions will be presented with enough background for the user to decide which is most reasonable for their specific situation. Background It is generally accepted that electrical current through the body is what does the damage. Levels of 9 to 22 milliamperes (mA) of current through a mans arm can prevent him from opening his hand [32]. This is known as the let go level. This current level causes an involuntary contraction of the muscles that cannot be voluntarily overcome. Table 7-1 lists different current levels and their effects on people.

7-4

Personnel Safety Considerations

Table 7-1 Reactions to Various Levels of Current [33] Threshold of Sensation This varies with skin chemistry; a few individuals may be able to discern 1/2 milliampere, while others are unable to discern levels up to 3 milliamperes. Mild Shock may startnot painful Muscular Contraction, But Can Let Go If not grasping an object, involuntary muscle contraction can cause the body to jerk away from the object. Painful Shock may start GFCI Breaker Will Trip Inability to Let Go (If grasping an object) The let-go current level is generally greater for larger persons; with adult males running in the upper portion of this range, and children running in the lower portion; the letgo current for infants may be as low as 3 milliamperes. Painful shock. Serious burns may begin, if contact time is long enough. Threshold of Unconsciousness and Possible Asphyxia Difficulty in breathing begins in the early portion of this range; at the end of the range, breathing is very difficult and suffocation may have begun. Fibrillation and Asphyxia may begin Fibrillation of the heart interrupts or reduces blood flow and exacerbates breathing difficulties and reduces brain function. Electric Toothbrush Will Operate (10 Watts) Fibrillation to Cardiac Arrest Above 200 ma, chest muscles may clamp heart and stop it during duration of shock, thus preventing fibrillation Severe Burns and Hemorrhage Severe muscle contractions occur. Depending upon contact pressure and moisture content of flesh at contact point, heat will turn moisture into steam causing an explosive expulsion of enough material to break contact. 40 Watt Light Bulb Will Operate

1 milliampere (0.001 ampere) 2 milliamperes 2 to 10 milliamperes 4 milliamperes 5 milliamperes 6 to 25 milliamperes @ 60 cycles a-c, 99.5% of women can let go with 6 ma; 99.5% of men can let go with 9 ma. (for direct current: 41 ma & 62 ma respectively)

20 to 50 milliamperes

40 to 100 milliamperes

90 milliamperes 100 milliamperes to 1-4 amperes

200-300 milliamperes

350 milliamperes

Note: Fibrillation current is sometimes quoted as starting at 20 to 40 mA. That value is for current through the heart muscles. This handbook addresses current through the body (such as hand-to-hand) of which only a portion passes through the heart.

While it is the current that does the damage, the current is very dependent on the source voltage, source impedance, the contact impedance, and the impedance of the person. Variations in impedance often make it impractical to set a Standard that specifies a current limit. As a result, most electrical safety Standards set voltage limits that are intended to achieve the desired levels of safety. The question of levels of electrical safety can be broken down into groups of people and situations, and then the effects can be better evaluated. For example: safety levels for trained 7-5

Personnel Safety Considerations

workers might be different from those applied to the general public. The levels that are safe for steady-state exposure will be different from the levels for very short exposures. Perhaps the only thing we can assume in this handbook is that the frequency of the electrical energy of concern will be primarily 50 Hz or 60 Hz. Duration of Exposure The effect electrical energy has on a person can be dependent on the duration of the exposure. A level of current that can cause ventricular fibrillation if applied continuously (steady-state), might not be a danger if it lasted 0.083 seconds (5 cycles at 60 Hz). Also, the level of ac interference from power systems into railroad facilities varies in amplitude and duration depending on the source and the path of the energy. For these reasons, two separate criteria are usually specified. The first is the steady-state level. Steady-state interference is ac energy that is either continuously present, or is present for long periods at a time (possibly intermittently). What constitutes a long period of time depends on the situation. The second criterion is for short duration exposures. Sometimes these limits are stated as being applicable for power systems with high speed relaying. High speed relaying refers to systems that act like circuit breakers and turn off the line if a fault occurs. Typically, these systems operate within five cycles, or 0.083 seconds at 60 Hz. Steady-State Exposure Electrical engineers understand transients and steady-state conditions. Transients are generally disturbances with a duration of less than a few cycles. (There are actually seven different definitions of transient in the IEEE Standard Dictionary of Electrical and Electronic Terms [34]) We will not be using the word transient much in this handbook, because ac interference from power line faults lasts longer than a few cycles and is, therefore, not a transient. In fact, the effect of a power line fault could more accurately be described as a steady-state condition of short duration. While there are transients that affect railroad signal systems and personnel, they are generally not caused by power systems. Things like lightning cause them. There are true transients in the power system, but they usually dont affect railroads because their levels are too low to damage equipment or create a shock hazard, and they are too short to have a noticeable affect on signal operation. So, what does steady-state really mean in this context? We will typically use the term steadystate to refer to conditions that are maintained for 2 seconds, or more. There are two reasons for this choice. First, when railroad signal equipment operation is an issue, durations greater than 2 seconds usually affect operation. Shorter duration events often do not affect operation. Second, the effect on the human body of current lasting 2 or more seconds approaches the effect of continuous exposure. While the ability of the human body to withstand currents lasting less than 2 seconds is dependent on the actual duration. 7-6

Personnel Safety Considerations

Now that we understand what we mean by steady-state, there is another term that we need to know: intermittent. Steady-state does not always mean continuous. The interference level that caused you problems an hour ago might be gone by the time you get your test equipment hooked up. Intermittent steady-state interference can be the hardest to track down. Short Duration Exposure The power system fault condition generally is of short duration, ranging from approximately 30 to 250 milliseconds for transmission systems. Reclosure3 after a ground fault is common in modern electric utility practice. Automatic reclosure can result in subsequent events starting within less than 0.5 seconds from the first. IEEE Standard 80 [35] notes that the cumulative effect of two or more closely spaced shocks has not been thoroughly evaluated, but a reasonable allowance can be made by using the sum of individual shock durations as the time of a single exposure. The induced voltage and currents for a single-phase to earth fault are typically much higher than for other fault conditions or the steady state. The summary provided by the IEC [36] consolidates available information into time-dependent current curves of increasing probability of ventricular fibrillation with no dependence on body weight. The threshold, where ventricular fibrillation is unlikely, is illustrated in Figure 7-1.
Table 7-2 Time/Current Zones for AC 15 Hz to 100 Hz (IEC 479-1: Table 4, 1994)

Refer to the definition in the Glossary, Chapter 18.

7-7

Personnel Safety Considerations

Figure 7-1 Time/Current Zones of Effects of AC Currents 15 Hz to 100 Hz (refer to Table 7-2 for instructions)(IEC 479-1: Figure 14, 1994)

The guideline that is most often used in the United States and Canada to assess the personnel shock hazard associated with fault current energization of conductors is IEEE Standard 80. This standard was developed to provide safety guidelines for persons working within ac power substations. The document provides guidelines for personnel safety for short duration shocks such as may occur during a transmission-line fault event. The guideline is primarily based upon the work of Dalziel, which extrapolates animal study results to humans. The criterion for safety is the current that will not cause ventricular fibrillation for 99.5% of the population. The current for 99.5% survival of a shock event is shown in Figure 7-2 for persons of two different weights (see the two straight lines in Figure 7-2). The survival current is a function of the shock duration. The equations presented in the IEEE standard are also shown on the graph. Evaluation of the current that will flow through a person that contacts an energized structure can be complex. The difficulty is caused by several unknown resistances in the current path from the object, through the person, and to ground or to a return conductor.

7-8

Personnel Safety Considerations

1.0

99.5 % Non-Fibrillation Current (Amps rms)

70kg (155lb) Person 0.5 Ib=0.157/t

0.1

50kg (110lb) Person 0.5 Ib=0.116/t

From IEC 479-1, 1994

0.0 0.01 0.1 1 10

Current Duration (seconds)

Figure 7-2 IEEE Std 80-1986 Guideline and IEC 479-1, 1994 - Survival Body Current

As noted above, significant difference in opinion exists regarding the applicability of bodyweight scaling for evaluating fibrillation thresholds. The IEC guidelines do not consider body-weight scaling, while the IEEE guidelines do (see the curved line in Figure 7-2). Philosophy of Setting Standards Traditionally, Standards makers have tried to find a balance between laboratory or academic evaluations, and actual accident records, in determining threshold levels and limits in standards. The steady-state limit for ac exposure is a good example. On the academic side, it is reasonable to use a body impedance value of 2000 Ohms when voltage is about 50 V rms ac. It is also reasonable to use 10 milliamperes as a let-go limit. Furthermore, it is reasonable to assume zero contact resistance, and that the source impedance is insignificant (<< 2000 Ohms). Applying Ohms Law to these reasonable assumptions results in a voltage threshold of 20 Volts rms ac. On the other hand, the following is from a 1973 IEC document [37]: 7. Experience with voltages not exceeding 50 V rms ac or 75 V dc From the replies given by several countries to a questionnaire, it appeared that there is no conclusive evidence in any of those countries of accidents occurring under usual circumstances at supply voltages not exceeding 50 V rms ac or 75 V dc and caused by a current passing through the body that led to serious injury. 7-9

Personnel Safety Considerations

To the best of our knowledge, that 1973 statement is still accurate today. Together, the laboratory/academic evaluations and real-world data make up the body of evidence upon which we base Standards. Such a body of evidence has usually led to establishing 50 V rms ac (or 60 V) as the limit for exposure to steady-state voltages. Standards citing a 50 V limit include: OSHA AAR/EEI Bluebook (60 V rms ac) IEEE 80 NEC NESC AREMA

But, there are exceptions. One of those exceptions can best be summarized by quoting from the Foreword of NACE Standard RP0177-95 [38]: Some controversy has arisen in the latest issue of this standard regarding the shock hazard stated in Section 5, Paragraph 5.2.1.1 and elsewhere in this standard. The reason for a more conservative value in this revision of the standard is that early work by George Bodier at Columbia University and by other investigators has shown that the average hand-to-hand or hand-to-foot resistance for an adult male human body can range between 600 ohms and (1) 10,000 ohms. A reasonable safe value for the purpose of estimating body currents is 1,500 ohms hand-to-hand or hand-to-foot. In other work by K.S. Gelges and C.F. Dalziel on muscular contraction, the inability to release contact would occur in the range of 6 to 20 milliamperes for adult males.(2) Ten milliamperes hand-to-hand or hand-to-foot is generally established as the absolute maximum safe let-go current. Conservative design would use an even lower value. Fifteen volts ac impressed across a 1,500-ohm load would yield a current flow of 10 milliamperes, thus the criterion within this standard is set at 15 volts. Prudent design would suggest an even lower value under certain circumstances. George Bodier, Bulletin de la Societe Francaise Des Electriciens, October 1947. C.F. Dalziel, The Effects of Electrical Shock on Man, Transactions on Medical Electronics, PGME-5, Institute of Radio Engineers, 1956. Similarly, the Canadian Standards Association (CSA) [39] also identifies a 15-volt value for pipeline induced voltage personnel safety. As explanation, they note: The value of 15 V has been selected as a practical mitigation level [emphasis added] that falls within generally accepted guidelines for exposure of the general public to continuous 60 Hz rms voltage. Some industries may accept different voltage levels where trained personnel or other technical factors are involved. Clearly, NACE and other pipeline Standards organizations have chosen to apply the calculated values despite the absence of accidents at these voltages. This process is not wrong. But the resulting value is not the only reasonable value.

7-10

Personnel Safety Considerations

There are differences between pipelines and other receptors of ac interference. With a pipeline, mitigation of ac interference is relatively straightforward. If you expect that voltages will exceed acceptable levels (steady-state or fault), you bury a bare metallic ribbon near the pipe, and bond that ribbon to the pipe at regular intervals. This practical mitigation eliminates the touch potentials by providing a local ground reference for the pipe. This procedure has a cost, but that cost is substantially less than the cost of the pipeline itself. With a railroad the situation is very different. You cannot ground the track, because the track must carry electrical signals. So, how do you mitigate unacceptable voltages levels? There are many ways (see Chapter 13). But all of these ways have their limits. None can eliminate excess voltage as effectively as grounding, in the way that pipelines do. When the voltage limit is very low, the cost of further mitigation can actually exceed the cost of the power line itself! Pipeline-type levels cannot be achieved with practical mitigation on railroads. So the difference is simply this: the pipeline industry can set an extremely conservative voltage limit and build to that limit with relative ease. But the railroad industry cannot. The cost of meeting such a limit would, in many cases, be unreasonable. In some cases it would simply be impossible. So the railroad industry needs to take care in setting standards. It would not be reasonable to place prohibitive restrictions on railroads or power companies especially since there is insufficient data of actual accidents to support such low voltage limits.
Usual Circumstances

One aspect of the philosophy of setting standards is what constitutes usual circumstances, and whether or not the same limits need to be imposed on both usual circumstances and unusual circumstances. Take another look at the quote from the IEC document: From the replies given by several countries to a questionnaire, it appeared that there is no conclusive evidence in any of those countries of accidents occurring under usual circumstances [emphasis added] at supply voltages not exceeding 50 V rms ac or 75 V dc and caused by a current passing through the body that led to serious injury. It is implied that accidents occurring under unusual circumstances, if any, would have been excluded from this survey. So we must define what constitutes usual or unusual circumstances before setting standards for these distinct cases. Historically it has been common practice to limit the scope of standards to usual circumstances. However, in creating standards some consideration should be given to unusual circumstances. Some examples of circumstances that may or may not be considered usual are: Pedestrians trespassing on railroad tracks (as opposed to pedestrians crossing the tracks at grade crossings) Vandals damaging or stealing wire, cable, or other conductors Failed railroad insulated joints

7-11

Personnel Safety Considerations

Power line contingencies (abnormal steady-state operation of power lines at higher loads due to equipment failure or maintenance - lasting minutes, hours, or days)

When planning a new power line parallel with a railroad, the computer modeling is based on worst case situations. But what constitutes a worst-case situation is sometimes elusive. One example of an area of disagreement is failed insulated joints. The question is: Is a failed insulated joint a normal (i.e., usual) condition? Most railroad signaling systems are designed to continue to work if only one insulated joint is shorted. (If a second joint fails, the signals fail.) But where magnetic induction causes ac interference, a single failed insulated joint can double the rail-to-ground voltage, increase the voltage across adjacent insulated joints by 50%, and increase rail-to-rail voltage ten-fold. To mitigate ac interference so that a railroad signaling system will continue to operate correctly with a failed insulated joint is often impossible. To mitigate ac interference so that personnel safety limits are maintained with a failed insulated joint can result in significant expense. If a failed insulated joint is not considered a usual circumstance, then a process must exist to inspect and repair them. In essence, this process becomes part of the mitigation. On the other hand, a balance is sometimes achieved by realizing that it is extremely unlikely that two or more unusual circumstances will happen at the same time and place. If every worstcase condition possible were assumed to be happening at the same time and place, then no amount of mitigation could guarantee safety. Sound engineering judgment is needed to evaluate these issues. Steady-State Limits As described above, steady state limits of ac interference for personnel safety have been debated for some time. In this section we will examine, in greater depth, some of the justifications used for the various values proposed. The current-based thresholds reviewed above reflect the present state of knowledge and are the most effective indicators, since the physiological response appears to be current or currentdensity dependent. Transforming these current-based thresholds into their corresponding voltagebased thresholds adds an additional level of complexity, uncertainty, and variability that can be contributed to by: The impedance of the current path through the body The impedance of the energized conductor circuit The type of coupling to the conductor from the power system

These factors suggest that care is essential in interpreting a voltage-based guideline or objective. The following provides a brief review of the above factors.
Body/Contact Impedance

The body current that results from an ideal voltage source can depend on several factors including: 7-12

Personnel Safety Considerations

the path through the body the conductor/skin contact area the pressure of the conductor/skin contact the moisture content and surface wetness of the skin.

The impedance to current flow through the human body can vary over a wide range, with the interface between the conductor and the skin playing a dominant role. The total body impedance depends on the voltage for low-impedance low-voltage sources, with the body impedance increasing for lower values of voltage. IEC 479-1 presents measured values for the total body impedance and its dependence on voltage for a hand-to-hand current path for a large contact area. Their impedance data, which is not exceeded by 5% of the population, is shown in Figure 7-3. As contact area is reduced, the impedance increases for a given voltage. For example, IEC 479-1 notes that the dry contact impedance for an area of 1000mm2 (about one 2 finger) is approximately an order of magnitude higher than for 8000mm (about the size of the palm of a hand). And, the impedance for an area of 100mm2 (about the tip of a finger) is an order of magnitude higher than for an area of 1000mm2.
10000
Percentage population with resistance

95%
Total Body Impedance (ohms)

50% 5% 1000

Large area contact: 8000mm2 (12.4 in2), Dry skin. 100 10 100
Touch Voltage (Volts rms)

1000

Figure 7-3 Values of Total Body Impedance Hand-Hand or Hand-Foot for AC 50/60 Hz (Trend Lines Through Data Values Presented in IEC 479-1)

The contact area and skin moisture conditions for common field maintenance situations in which railroad system conductors could be energized by power line system coupling cannot readily be determined. However, the IEC results of Figure 7-3 should provide a conservative range of total body impedance for use in identifying railroad safety objectives for lower voltage, large contactarea exposures. For example, the contact area for railroad track maintenance may be expected to be larger than for contact with typical communications/signal conductors. Induced rail voltage 7-13

Personnel Safety Considerations

can also transfer to a locomotive or rail car on an energized segment of track, resulting in a voltage between grab bars (a.k.a. hand holds and handrails) and the earth near the track. Safe contact voltage, for either the rails or the rolling stock, may be relatively low due to the larger contact area involved. The IEC has created model circuits that simulate the body impedance for evaluating leakage current or touch current over a broad range of frequencies. Test instruments based on these circuits have been constructed and are used to simulate the electrical properties of a human body. At 60 Hz the resistance of these instruments is 2000 ohms [40], to simulate the resistance for a person grasping a handle or other grippable conductor. Similarly, a UL leakage-current instrument has a resistance of 1500 ohms at 60 Hz.
Conductor Circuit Impedance

The impedance of the energized conductor circuit can have a significant effect on the current that will flow through a person who contacts the conductor. The equivalent source impedance of railroad-signal conductors energized by 60-Hz power coupling can range from less than one ohm for rails, to thousands of ohms for signal pole-line conductors. Thus, measuring or calculating the voltage of the conductor for normal conditions does not provide a good indication of the shock current that can flow through a person who touches the conductor. The current that flows through a person contacting an energized conductor depends not only on the voltage but also on the resistance of the person and the impedance of the conductor circuit.
Source Impedance

The voltage and the current that can be drawn from the conductor will be influenced by the coupling mechanism (electric induction, magnetic induction, or earth conduction) between the power system and a signal-system conductor. Electric induction is typically only important for high-impedance receptor circuits like pole-line signal circuits, while magnetic induction or earth-current coupling is generally more important for low-impedance circuits such as rails. Electric Induction Excitation Electric induction (capacitive coupling) results in a distributed current being collected along the exposed length of a conductor as is shown in Figure 7-4. The figure also shows an equivalent circuit for the arrangement, in which the total current collected by electric field coupling to the conductor is represented by a current source, I. Before the person contacts the energized conductor, as represented in the equivalent circuit with an open switch in series with the resistance of the person, all the collected current flows through the parasitic wire-to-ground capacitive impedance. Signal conductors that are isolated from earth, such as many pole-line signal circuits, can have relatively high values of capacitive impedance to earth, which can result in high values of steadystate electrically induced voltage. The resistance through a person can be much less than the high impedance of the conductor circuit. For that condition, the current that flows through a person contacting the conductor (with the switch closed in Figure 7-4b) will not be sensitive to the impedance of the person. The hazard associated with this type of circuit should not be evaluated

7-14

Personnel Safety Considerations

by the voltage induced onto the circuit, Voc, but by the current, Iman, that can flow through a contacting person4.

dis cur tribute ren d t

dis cur tribute ren d t

(a) Electrostatic-Coupled Current to Person

Cwg

Voc Rman I man

(b) Equivalent Circuit of Energized Conductor and Man


Figure 7-4 Electrostatic Coupling to Conductor Touched by Man

An example of the effect of electric induction would be a large truck parked under a transmission line. The open circuit voltage measured from the truck to ground may be thousands of volts. But, if an electrical connection is made to ground (sometimes through a resistor to simulate a person), the current to ground may only be a few milliamperes. The NESC requires transmission lines be designed such that this short-circuit current to ground is no greater than 5 mA for the largest anticipated vehicle.

7-15

Personnel Safety Considerations

Magnetic Induction Excitation Magnetic coupling results in an induced voltage that is distributed along the exposed length of a conductor. Signal conductors that are well isolated from earth, such as many pole-line conductors, may have impedance high enough to prohibit hazardous magnetically induced current from flowing through a person contacting the conductor. If an abnormal condition develops on a normally high-impedance circuit, such as low impedance to ground at a remote location, the circuit is no longer high impedance. The isolated signal-line condition is in contrast to telephone or communications conductors that are earthed at some remote location (central office). Rails may have a sufficiently low impedance to earth such that the current through a person contacting the rail is controlled primarily by the impedance of the person and the voltage of the rail. The hazard associated with this type of circuit should be evaluated based on the voltage induced onto the circuit, Voc5.
Voltage Limits Derived From Current Effects

For high impedance receptor circuits excited by electric induction, the short-circuit current to ground is a good measure of the hazard. But for low impedance receptor circuits excited by magnetic induction or conduction, the victim and the circumstances limit the current. So, a direct measurement of the current is impractical. The common practice is to select a value of opencircuit voltage, Voc, as the maximum acceptable limit for the situation being evaluated. As described earlier, there are various philosophies and methods for arriving at voltage limits. Typically, tests are conducted under various conditions, statistical distributions are identified, safety factors are applied, and the results are compared to recorded experience. If it all fits, those results become the limit. Unfortunately, test results usually vary wildly depending on conditions. People also have very different ideas as to what safety factors are appropriate, and the recorded experience doesnt often line up with the theoretical result. In addition to these issues, different industries have different concerns that impact the selection of an open-circuit voltage limit.
Other Considerations

As noted earlier, a variety of standards exist that employ a 50-Volt maximum limit on steadystate 60 Hz ac rms exposure. These guidelines are applied for a wide range of conditions. But, it is important to apply good engineering judgment to specific situations. The following is provided for the reader who wants to delve deeper into the topic: The pipeline industries in the United States [41] and Canada [42] have identified a steady-state personnel safety touch potential guideline of 15 volts for power-line-induced voltage on pipelines. This guideline is consistent with a 10 mA let-go threshold and a 1500 ohm body-path impedance, with a low impedance electrified object.

If a circuit is excited by magnetic induction the voltage will usually be less than 50 Volts. If a short circuit is created to ground, the current to ground may be many Amperes because of the low impedance of the circuit. In fact, the resistance of the ground rod will probably be the limiting factor in the current to ground. The 5 mA limit for electric induction does not apply for circuits excited by magnetic induction.

7-16

Personnel Safety Considerations

This is in contrast to the telephone-industry guidelines that identify considerably higher safe touch potentials. The IEEE Std 776 [43] identifies that 50 V rms continuously induced with respect to ground at 60 Hz on telecommunications facilities has historically been considered an upper threshold by many telecommunications companies in North America for personnel safety. Similarly, the International Telegraph and Telephone Consultative Committee (CCITT) [44] notes that To avoid danger, it is recommended that the permissible continuous induced voltages be limited to 60 volts rms. This applies to screened or unscreened cables or open wire lines to which access is required for work operations by staff. The Canadian Standard CSA-C22.3 No.3 identifies an acceptable longitudinally induced voltage in railway signaling and communications circuits as 50 V ac rms under normal power line conditions. Under special conditions, 150 V is acceptable, which may require special instructions and marking. This Canadian standard also limits the voltage across insulated rail joints to 50 V, and limits the maximum rail-to-remoteearth voltage to 25 V. One difference between the pipeline and telephone industry guidelines is the size of the likely contact area to an energized conductor. The energized conductor for the pipeline industry is assumed to be of large diameter, such that it can be contacted with a large area of the hand. In contrast, the small conductors associated with communication/signal cable and open-wire 2 2 conductors have a small surface contact area, perhaps 100 mm (0.155 in = 0.39 in * 0.39 in) or less. The small contact area may cause total body impedance in the range of 200,000 ohms for dry hands at 25 volts [45]. Thus, for a given conductor voltage, the body current for contact with small-diameter signal and communication conductors can be much less than for large conductors such as pipes, for a comparable conductor voltage (all other factors being equal). The IEC [46] defines a grippable part as a part of the equipment which could supply current through the human hand to cause muscular contraction round the part and an inability to let go. Parts which are intended to be gripped with the entire hand are assumed to be grippable... Why does it matter if a part is grippable? This section is concerned with steady-state exposure. As such, the common limit is a let-go level. But, for let-go to be a concern, the object must be grippable able to get your fingers around it and hold on. If it is too large to be grippable, like a 12-inch diameter pipe, then let-go might not be an issue. (Careful here: The pipe has a valve somewhere and that valve has a grippable handle.) This definition of a grippable conductor appears to be applicable to un-insulated ends of track lead wires, and possibly to rails, as well as rail cars or locomotives, to the extent that the human hand can get around the part. For many years the idea of applying probabilistic risk assessment has been discussed. The idea involves applying a rigorous process to determine the likelihood of the worst-case conditions and the appropriateness of the safety factors applied in selecting safety limits. Work in this area is just beginning. So it is necessary to consider the specific situation in determining a voltage limit. For the steadystate exposure presented in this section (greater than 2 seconds), the let-go criteria is the usual standard. But, selecting the appropriate let-go current value and arriving at a voltage target involves many considerations. Some of them are: 7-17

Personnel Safety Considerations

General public or trained personnel Adults or children Grippable or not grippable Hand contact or step potential Boots or bare feet Dry skin or perspiration Dry earth or salty slush

Possible Target Values

The bad news is that this Handbook will not give you one clear design limit. There are too many case-specific variables to do that with due diligence, as regards both safety and practical reality. What follows are a few possible steady-state levels and the assumptions made to derive them: 50 V ac rms No conclusive evidence of accidents under usual circumstances that led to serious injury below 50 V. Little or no statistical data.

15 V ac rms 10 mA let-go current 99.3% of range for men 64% of range for women 93.2% of range for all adults Large adult hands (8,000 mm ) on cylindrical electrodes (both hands) Valid for 95% of the adult population Between fresh water and conductive solution levels
6 2

1500 Ohm body path resistance

In other words, this value is unnecessarily low for a situation if any of the following are true: Contact area is less than both hands firmly grasping conductors The individual in question is one of the majority of adults who have a let-go threshold above 10 mA

From IEC 479-1: Table 1 shows 5% total body impedance for 25 V is 1750 . Extrapolated to 15 V would be about 2000 (Figure 6 shows the 5% total body impedance for 10 V to be about 3500 ). From page 19, wetted with fresh water are 10% to 25% lower (1800 to 1500 ). And, with conductive solutions, down to half the values measured in dry conditions. (1800 to 1000 ).

7-18

Personnel Safety Considerations

The individual in question is one of the 95% of adults with higher total body impedance Dry conditions exist (skin, soil, etc.) Wet with clean water conditions (as opposed to conductive solutions) The individual is a child the contact area of hands would be too small

But what if the 50% values are used for total body impedance and let-go? 2625 (IEC 479-1) x 15 mA [47] = 39.4 V This is less than 50 V. So why isnt half the population stuck to energized metal things? For this criteria to apply all the following must happen: Elevated voltage must be present (>39.4 V) Adult-sized hands for contact area Grasp objects with both hands (or be standing barefoot in salt water) Have hands saturated with salt water So, what operating conditions do we design for? The common practice is to design for a worst case condition. That usually includes the following: Maximum loading on power line Single-circuit operation of power lines (phase cancellation eliminated) Maximum point-to-point voltage (usually across an insulated joint on the rails)7

Common practice does not include the following: Failed insulated joints Grounded track circuits (shorted arresters or other inadvertent grounds) Defective power system or railroad equipment

The bottom line question seems to be whether the 50 V ac rms safety limit is adequately safe. The reason for doubt comes from the fact that it is possible for a person to exceed the let-go current level at lower voltages if they are very well grounded at one point, and very well connected to an energized conductor at another. But there are several reasons that this is unlikely:
7

Most power transmission lines are rarely operated above half of their rated capacity. Multi-circuit power lines use phase cancellation during normal operation to reduce induction.

Common practice used to be to design to maximum rail-to-ground voltage of 50 V ac rms. In cases of magnetic induction, the IJ voltage can be twice the rail-to-ground voltage. So, this change alone cut design levels in half.

7-19

Personnel Safety Considerations

Persons contacting the rails are standing on crushed rock ballast that is a very poor path to ground. Persons contacting equipment enclosures are protected by the buried ground wire network in the earth around the enclosure. Persons contacting the signal wires have a much smaller contact surface area and greatly increased total body resistance. With point-to-point voltages limited to 50 V, rail-to-ground voltages will rarely exceed 25 V. With only one or two IJs per mile (typically) it is very unlikely that a person would contact the rails directly at the point of greatest exposure. For signal personnel who might be exposed to leads across an insulated joint, the contact area is so small that the total body impedance would limit the current.

We are benefiting from the fact that a hazardous combination of events is improbable. But this combination of events is not impossible. If circumstances arise where several of these factors are present at once, it might be prudent to limit risk by removing the personnel from the area, establishing work rules to remove the risk, or taking action to reduce the voltage present. While the various 50 Volt ac rms standards for steady-state exposure apply in most areas (including North America), special cases may arise requiring additional restrictions. Limits for Faults and other Short-Duration Events A reasonable personnel safety objective for power-system coupling to railroad-system conductors is to minimize the chance for lethal body current for fault-conditions. For exposure to ac current from short-duration events, ventricular fibrillation is the primary concern. Consideration should be given to who will be contacting the conductors. For example, informed maintenance personnel, using special procedures, may tolerate higher levels of available current than maintenance personnel who expect a non-energized circuit with no special precautions, or if the general public has ready access to the conductor. Selection of more than one objective may be warranted if different levels of personnel awareness and precaution are incorporated. For short-duration current to the human body, such as for contact with a power-fault energized conductor, the recognized United States standard is IEEE Std-80, which was developed to provide safety guidelines for persons working within ac power substations. The IEEE has standardized on a simple model to relate the voltage on a fault-energized object to the current that may flow through a person contacting the object. The model, which is shown in Figure 7-5, assumes the resistance of the person to be 1000 ohms and ignores any contact resistance from person to object. The model includes the contact resistance from two feet to soil but ignores any shoe resistance. The feet contact to the soil is considered equivalent to two 6-inch diameter metal disks, with mutual impedance ignored.

7-20

Personnel Safety Considerations

Ib V

1000 ohm

Soil

Ib =
Figure 7-5 IEEE-80 Fault Touch Potential Model

V (1000 + 1.5 )

The model of Figure 7-5 allows the safe touch potential to be determined that will cause the 99.5% non-fibrillation body current of Figure 7-2. The 99.5% safe touch potential is plotted in Figure 7-6 for a 70-kilogram (155 lb) person and in Figure 7-7 for a 50 kg (110 lb) person, for several values of soil resistivity. For example, for a soil resistivity of 100 ohm-m (10,000 ohmcm), and a fault clearing time of 0.1 second (6 cycles), the curves of Figure 10 give the safe touch potential as approximately 570 volts. Use of the values for a smaller mass person may be more appropriate in situations where access by the general public to an energized conductor is likely. While the IEEE-80 guideline for personnel safety for power faults depends on both the body weight and fault duration, other guidelines such as the CCITT [48] and the Canadian Standards Association [49] do not depend directly on those parameters. These two guidelines instead recommend a voltage limit of 430 volts rms, for a fault on a line that is constructed to usually accepted technical standards for the class of line involved. The tolerable voltage is increased to 650 volts for a fault on a high-reliability line. A high reliability line is characterized by the duration of fault events: faults never exceeding 0.5 seconds duration, and are less than 0.2 seconds for the majority of cases.

7-21

Personnel Safety Considerations


10000

70 kg (155 lb) person

Safe Touch Potential (volts)

Soil Resistivity (ohm-m)


1000

1000 500 100

100 0.01

0.1

Shock Duration (seconds)

Figure 7-6 IEEE Std 80-1986 Guideline for 99.5% Safe Touch Potential (70 Kg Person)
10000

50 kg (110 lb) person

Safe Touch Potential (volts)

Soil Resistivity (ohm-m)


1000

1000 500 100

100 0.01

0.1

Shock Duration (seconds)

Figure 7-7 IEEE Std 80-1986 Guideline for 99.5% Safe Touch Potential (50 Kg Person)

7-22

Personnel Safety Considerations

Possible Target Values

Computer modeling can be used to predict fault current magnetic induction into railroad systems. The same computer models can be employed to evaluate various mitigation options. But, the target voltage values for safety under fault current induction need to be predetermined according to the site-specific situation. The 430 V rms and 650 V rms levels used by CCITT, the Canadian Standards Association, and the 1977 AAR/EEI Bluebook can be used, provided that the levels can be reasonably met with practical mitigation, and there are no unusual circumstances requiring lower levels. In any case, it is reasonable to evaluate the situations using the IEC 479-1 method or the IEEE Std-80 method (or both) to insure adequate safety and to insure mitigation is not unnecessarily expensive. The worst-case fault is modeled to determine the voltage. What constitutes the worst-case includes: Modeling various fault locations to identify the one which produces the maximum voltage Fault the closest phase conductor to the track under elevated temperature final sag conditions Track circuit lightning arrester clamping ability may or may not be included in the worst case conditions by mutual agreement between the railroad and the power company.

Electric and Magnetic Field Standards and Guidelines Introduction to Standards and Guidelines In 2003, there were no U.S. federal government health standards or guidelines related to powerfrequency electric and magnetic fields. Two U.S. organizations have developed health guidelines for occupational and public exposure to power-frequency electric and magnetic fields. Some U.S. states have electric or magnetic field limits for certain voltage classifications of transmission lines within the right-of-way or at the edge of the right-of-way. These standards were established to limit electrical effects rather than being created from a health risk perspective. Standards or guidelines for exposure to electric and magnetic fields exist in some foreign countries. Some countries have simply adopted or modified international organizational guidelines while others seem to have been developed for a particular occupation (such as welders). Many other states and countries have considered setting magnetic field standards. Sweden has established standards related to computers and computer monitors. State Standards and Recommendations Related to Transmission Lines There are at least six states in the U.S. that have adopted engineering-based guidelines or standards for transmission line electric fields; two of these states also have standards for magnetic fields. The purpose of most of these standards is to make the field levels from new power lines similar to the field levels from existing lines or to minimize the potential for spark discharge from the induced current on large vehicles in the electric fields of 345-765 kV transmission lines. Table 7-3 presents a summary of these standards [50]. 7-23

Personnel Safety Considerations Table 7-3 State Regulations that Limit Field Strengths on Transmission Line Rights-of-Way Electric Field Limit State Within the Right-of-Way 8 kV/m for 69 230 kV lines Edge of Rightof-Way 3 kV/m for 69 230 kV lines Magnetic Field Limit Within the Right-of-Way ------Edge of Rightof-Way 150 mG for 69 230 kV lines (max load)

Florida

10 kV/m for 500 kV lines

2 kV/m for 500 kV lines

200 mG for 500 kV lines (max load)

250 mG for double circuit 500 kV lines (max load) Minnesota Montana 8 kV/m 7 kV/m maximum for highway crossings New Jersey New York ------11.8 kV/m 11 kV/m for private road crossings 3 kV/m 1.6 kV/m ------------------200 mG (max load) ------1 kV/m -------------------------

7 kV/m for highway crossings North Dakota Oregon 9 kV/m 9 kV/m -------------------------------------

7-24

Personnel Safety Considerations

In 1989, the California State Department of Education released a school site selection and approval guide [51] that contains recommendations related to electric and magnetic fields. The School Facilities Planning Division has established the following limits for locating school sites near certain high voltage power transmission line easements: 100 feet from edge of easement for 100 - 110 kV power line 150 feet from edge of easement for 220 - 230 kV power line 250 feet from edge of easement for 345 kV power line

These limits are based on an electric field strength graph developed by EPRI, published in Background on Electromagnetic Fields and Human Health [52]. The guide also recommends that the local electric utility be contacted to determine if any additional power lines or upgrades to existing power lines are planned. The Institute of Electrical and Electronics Engineers (IEEE) developed a standard for measuring power-frequency electric and magnetic fields near power lines. This standard, initially developed in 1979, has been periodically revised and updated through March, 1995 [53]. The purpose of this standard is to establish uniform procedures for the measurement of power-frequency electric and magnetic fields from overhead power lines. The standard also describes procedures for the calibration of instrumentation used to conduct these measurements. In addition to the transmission line electric and magnetic field standard, the IEEE has also developed a guide for the measurement of quasi-static magnetic and electric fields (describing various types of survey characteristics, goals and measurement methods) [54]. Guidelines for Exposure to 50/60 Hz Electric and Magnetic Fields Two organizations have developed health guidelines for occupational and public exposure to electric and magnetic fields: the International Commission on Non-Ionizing Radiation Protection (ICNIRP) [55] and the American Conference of Governmental Industrial Hygienists (ACGIH) [56]. Tables 7-4 and 7-5 present a summary of the electric and magnetic field levels of these guidelines respectively. The ICNIRP established these guidelines to provide protection against known adverse health effects. While the ICNIRP reviewed all of the scientific literature, the only adverse effects on humans that were fully verified by a stringent evaluation were short term, immediate health consequences (such as nerve and muscle stimulation, shocks and burns, etc.) [57]. The ACGIH established threshold limit values to which it is believed that nearly all workers may be exposed repeatedly without adverse health effects, based upon an assessment of available data from laboratory research and human exposure studies. The threshold limit values were developed as a guideline to assist in the control of health and safety hazards [58].

7-25

Personnel Safety Considerations

Both the ICNIRP and ACGIH guidelines are based on established adverse health effects (such as burns, shocks, nerve stimulation, etc.). Electric and magnetic field levels as specified in these guidelines, and which would cause these types of effects, are much higher than typical levels found in residential and most occupational environments, and are summarized below.
Table 7-4 Summary of ICNIRP 50/60 Hz Exposure Guidelines International Commission on Non-Ionizing Radiation Protection Guidelines Exposure (60 Hz) Occupational Reference Levels for TimeVarying Fields Current Density for Head and Body General Public Reference Levels for TimeVarying Fields Current Density for Head and Body 4.167 kV/m (4,167 V/m) 2 mA/m2 (5 kV/m) 0.833 G (833 mG) 2 mA/m2 (1 G) 8.333 kV/m (8,333 V/m) 10 mA/m2 (25 kV/m) 4.167 G (4,167 mG) 10 mA/m2 (5 G) Electric Field Magnetic Field

Table 7-5 Summary of ACGIH 60 Hz Exposure Guidelines ACGIH Occupational Threshold Limit Values for Sub-Radio Frequency Fields Electric Field Occupational exposures should not exceed: 25 kV/m (from 0 Hz to 100 Hz) Magnetic Field Occupational exposures should not exceed: 60 Hz: 10 G (10,000 mG)

Prudence dictates the use of protective devices (e.g. suits, gloves, insulation) in fields above 15 kV/m.

50 Hz: 12 G (12,000 mG)

For workers with cardiac pacemakers, For workers with cardiac pacemakers, maintain exposure at or below 1 kV/m. the field should not exceed 1 G (1,000 mG).

7-26

Personnel Safety Considerations

International Standards Electric and magnetic field exposure guidelines have been established in some foreign countries and by some international organizations. These foreign standards vary in regulatory power, from guidelines and recommendations to proposed or existing standards to established regulations and orders. While many other countries have considered establishing magnetic field standards, there are no national or foreign regulations limiting exposure to power-frequency magnetic fields from power lines. Tables 7-6 and 7-7 present a summary of international electric and magnetic field standards [59].
Table 7-6 Summary of International Electric Field Standards kV/m Public Country Australia [60] Austria [61] Czechoslovakia [62] Germany BFE [63] Exposure area 1 Exposure area 2 Italy [64] Poland [65], [66] Switzerland [67], [68] UK NRPB [69] USA ACGIH [58] USSR [70] Organization CENELEC [71] CEU [72] IRPA [73] ICNIRP [57] 5 , 10 5
p a c e f g,h k

Occupational As IRPA 10 to 30 15 21.32, 30, 30 6.67


c b

Status G PS S O, R O, R O O G, O G
r

Basis C C P, H C C H P, H C C C P, H C, P C C C, P

As IRPA 5, 10
a

5 , 10
j

1 , 10 5 12

15, 20 12.3 12 25

G, R
m

5 to 25 10

O PS, R
n

10 to 30b 6.1, 12.3, 19.6 10 to 30 25


b

PD G G

All at 50 Hz except IRPA (50/60 Hz) and ACGIH (60 Hz). Where levels for 16 2/3 Hz are specified, they range from the same to three times those at 50 Hz. C : limitation of induced Current density G : Guideline or recommendation H : Health concern for possible effects O : Order, rule, regulation or decree, often with legal force P : Perception of spark discharges or tingling sensations PD : Proposed Directive regarding the exposure of workers to physical agents (annex IV) PS : Pre-Standard R : Reference or investigation levels may sometimes be exceeded S : Standard, sometimes with legal force a : for up to a few hours per day and can be exceeded for a few minutes (up to 20 kV/m for 5 minutes in Austria) per day provided precautions are taken to prevent indirect coupling effects b : depending on the duration (t, hours per work day) of exposure , t < 80/E for E between 10 and 30 kV/m, although the exact interpretation of this formula differs between the three standards which use it c : exposure area 1 (controlled areas or short-time exposure) 8, 2 and 1 hours/day respectively e : exposure area 2 (longer-time exposure or areas where fields are not normally expected)

7-27

Personnel Safety Considerations


f: in areas or environments in which it may reasonably be expected that members of the public will spend a significant part of the day g : in cases in which exposure may reasonably be assumed to be limited to a few hours per day h : minimum distances of buildings to overhead power lines are also specified j: 1 kV/m applies where there are homes, hospitals, schools and the like k : 2 hours maximum m : depending on the duration (t, hours per work day) of exposure, t = 50/E 2 for E between 5 and 20 kV/m; between 20 and 25 kV/m, only 10 minutes exposure is permitted n : various actions would have to be carried out or requirements met before exceeding each of these levels p : for up to 24 hours per day this restriction applies to open spaces in which members of the general public might reasonably be expected to spend a substantial part of the day, such as recreational areas, meeting grounds, and the like r : public Federal recommendation (legally binding ordinance being considered); occupational legally binding protection of workers

Table 7-7 Summary of International Magnetic Field Standards Gauss (rms) Public Country Australia [60] Austria [61] Germany BFE [63] Exposure area 1c Exposure area 2 Italy [64] Poland [74] Switzerland [67, 68] UK NRPB [69] USA ACGIH [58] USSR [75] Organization CENELEC [71] CEU [72] IRPA [73] ICNIRP [57] 1 , 10 1
n a e g h,j p s, d

Occupational As IRPA 5, 50
b, d

Status G PS O, R O, R O O G, O G
r

Basis C C C C H C C C C W

As IRPA 1, 10
a

13.6, 25.5, 42.4c, d 4.24 1 , 10 1 16 5 , 50 4 16 10 18 to 75 6.4, 100c


p k

G, R

16d 2, 4, 6.4 5 , 50 5
m b, d

PS, R PD G G

C C C C

All at 50 Hz except IRPA (50/60 Hz) and ACGIH (60 Hz). Where levels for 16 2/3 Hz are specified, they are three times those at 50 Hz. C: G: H: O: PD : PS : R: S: W: a: b: c: d: e: limitation of induced Current density Guideline or recommendation Health concern for possible effects Order, rule, regulation or decree, often with legal force Proposed Directive regarding the exposure of workers to physical agents (annex IV) Pre-Standard Reference or investigation levels may sometimes be exceeded Standard, sometimes with legal force these values seem to have been developed primarily for electric-arc Welding for up to a few hours per day and can be exceeded for a few minutes (up to 20 G for 5 minutes in Austria) per day provided precautions are taken to prevent indirect coupling effects (additional body currents in Austria) maximum exposure duration is 2 hours per work day exposure area 1 (controlled areas or short-time exposure) 8, 2 and 1 hours/day respectively higher values are given for limbs exposure area 2 (longer-time exposure or areas where fields are not normally expected)

7-28

Personnel Safety Considerations


g: h: j: k: m: n: p: r: s: in areas or environments in which it may reasonably be expected that members of the public will spend a significant part of the day in cases in which exposure may reasonably be assumed to be limited to a few hours per day minimum distances of buildings to overhead power lines are also specified depending on duration of exposure from 8 to 1 hours per work day various actions would have to be carried out or requirements met before exceeding each of these levels for up to 24 hours per day this restriction applies to open spaces in which members of the general public might reasonably be expected to spend a substantial part of the day, such as recreational areas, meeting grounds, and the like for whole working day public Federal recommendation (legally binding ordinance being considered); occupational legally binding protection of workers depending on duration (t, hours per work day) of exposure according to D = H2 t, where H is the field strength in kA/m and D = 1.28 (kA/m)2 h (this gives 8 hours at 5 G and 5 minutes at 50 G)

Swedish Standards for Computers and Monitors Presently, there are two electric and magnetic field measurement and emissions standards which have been developed in Sweden. These standards relate to electric and magnetic field emissions from computer monitors and systems. The earlier standard, referred to as MPR II, was developed by the Swedish Board for Technical Accreditation (SWEDAC) and provides measurement methods for computer monitor electric and magnetic field emissions within certain frequency ranges. This standard was established on the basis that the electric and magnetic fields around a computer monitor should not tangibly increase the field levels that are generally found in their immediate environment. A more recent standard, developed by the Swedish Confederation of Professional Employees (TCO), places more restrictive limits on computer monitor field emissions and also includes emissions from computer systems. The TCO standard, which was established in 1992 and expanded in 1995 and again in 1999, also includes guidelines related to energy consumption, screen flicker, luminance, keyboard use, and other compliance requirements. Table 7-8 presents a comparison of the differences between the MPR II and TCO standards regarding electric and magnetic field emissions. Around the monitor, the MPR II standard prescribes 3 levels of 16 measurement points each. These measurements are conducted at sixteen equidistant locations around the monitor (at the center of all four sides, at the four corners, and at the midpoint between these eight locations), at a distance of 50 cm (approximately 20 inches) away from the monitor. The TCO standard prescribes that measurements be taken at a distance of 30 cm (approximately 12 inches) in front of the monitor and at 50 cm (20 inches) along the sides of the monitor (except for Band II magnetic fields and the static field, which are measured at 50 cm (20 inches) in front of the monitor screen). Electric and magnetic field emission requirements for computer monitors have remained constant from the TCO92 standard through the TCO95 and TCO99 standards. The TCO95 standard also introduced new computer monitor requirements, such as: 1. the computer monitor to have an automatic power-down function (the display switches off automatically after a preset time of inactivity) 2. the manufacturer must declare and provide the energy consumption power requirements of the monitor 3. the monitor must conform to the European fire and electrical safety requirements 4. the manufacturer must have the monitor tested and certified by a TCO-approved testing facility as compliant 7-29

Personnel Safety Considerations

Although not directly related to electric and magnetic fields, the new TCO95 requirement for automatic power-down of the monitor after inactivity lowers field levels around the monitor by decreasing power consumption. Other non-EMF related issues were also addressed, including certification of the computer system and keyboard, ergonomic qualities, energy efficiency, and ecology. TCO99 standards continue this expansion by including flat screen displays, display image quality (luminance, contrast, flicker and reflection), power consumption, and ecology. Values for electric and magnetic field emissions are the same as for TCO92, but the demands for resistance to influence from external magnetic fields (VDT interference) are now addressed. Manufacturers of computer monitors and systems are now providing TCO certification documentation or labeling, to show compliance with either the TCO92, TCO95, or TCO99 requirements respectively.
Table 7-8 Comparison of MPR II and TCO Guidelines Electric Fields Frequency Range 0 Hz (static field) 5 Hz 2 kHz 2 kHz 400 kHz Magnetic Fields Frequency Range 5 Hz 2 kHz 2 kHz 400 kHz TCO Standard Less than or equal to 200 nT (2 mG)* Less than or equal to 25 nT (0.25 mG)* MPR II Standard Less than or equal to 250 nT (2.5 mG) Less than or equal to 25 nT (0.25 mG) TCO Standard Less than or equal to +/- 500 V Less than or equal to 10 V/m* Less than or equal to 1.0 V/m* MPR II Standard Less than or equal to +/- 500 V Less than or equal to 25 V/m Less than or equal to 2.5 V/m

* Note: The difference between the TCO and MPR II guidelines is in fact larger than shown in the table, since the TCO values are measured 30 cm (approximately 12 inches) in front of the monitor screen, compared to the distance of 50 cm (approximately 20 inches) prescribed by MPR II.

Safe Working Clearances from Power Lines Safe working clearances from power lines must be maintained. Various standards exist establishing safe working clearances from power lines. In the United States OSHA sets such standards. Always check with the local power company for their working clearances for untrained workers (their standard clearances may be greater than OSHAs). Remember that the trained workers with reduced clearance requirements are power company workers with current training. This normally doesnt include railroad signal personnel.

7-30

Personnel Safety Considerations

Figure 7-8 shows an example of a working clearance card distributed by an electric company. Figure 7-9 shows as example of a signal installation with marginal working clearance from the top platform. No one is sure how they managed to install this equipment.

Figure 7-8 Sample of a Working Clearances Card

Figure 7-9 Railroad Signal Equipment with Marginal Working Clearance

7-31

Personnel Safety Considerations

References
32. Let-Go Currents and Voltages, C. F. Dalziel, F. P. Massoglia, AIEE Paper 56-111, May 1956. 33. Practical Utility Safety by Allen L. Clapp, published by Clapp Research, Inc. (c) 1999. 34. IEEE Standard Dictionary of Electrical and Electronic Terms, IEEE Std 100-1996, December 1996. 35. ANSI/IEEE Std 80-1986, IEEE Guide for Safety in AC Substation Grounding. 36. IEC 479-1, Effects of current on human beings and livestock - Part 1: General Aspects, 1994. 37. Effects of Current Passing Through the Human Body, International Electrotechnical Commission, Committee No. 64, April 1973. 38. Mitigation of Alternating Current and Lightning Effects on Metallic Structures and Corrosion Control Systems, National Academe of Corrosion Engineers (NACE), Standard RP0177-95, Item No. 21021, March 1995. 39. Principles and Practices of Electrical Coordination Between Pipelines and Electric Supply Lines, CAN/CSA-C.22.3, No. 6-M91, Canadian Standards Association, 178 Rexdale Blvd, Etobicoke, Ontario, Canada M9W 1R3. 40. IEC 990 Methods of measurement of touch-current and protective conductor current, 1990. 41. Mitigation of Alternating Current and Lightning Effects on Metallic Structures and Corrosion Control Systems, Standard RP0177-95, National Association of Corrosion Engineers (NACE), P.O. Box 218340, Houston, TX 77218. 42. Principles and Practices of Electrical Coordination Between Pipelines and Electric Supply Lines, CAN/CSA-C.22.3, No. 6-M91, Canadian Standards Association, 178 Rexdale Blvd, Etobicoke, Ontario, Canada M9W 1R3. 43. IEEE Std 776-1992, IEEE Recommended Practice for Inductive Coordination of Electric Supply and Communications Lines, IEEE, Inc., 345 East 47th Street, New York, NY 10017-2394. 44. CCITT, Directives concerning the protection of telecommunication lines against harmful effects from electric power and electrified railway lines, Volume VI, Danger and Disturbance, Geneva 1989, ISBN 92-61-04041-3. 45. IEC 479-1, Effects of current on human beings and livestock - Part 1: General Aspects, 1994. 46. IEC 990 Methods of measurement of touch-current and protective conductor current, 1990. 47. Let-Go Currents and Voltages, C.F. Dalziel and F.P. Massoglia, AIEE Winter General Meeting, 1956. 48. CCITT, Directives concerning the protection of telecommunication lines against harmful effects from electric power and electrified railway lines, Volume VI, Danger and Disturbance, Geneva 1989, ISBN 92-61-04041-3.

7-32

Personnel Safety Considerations

49. Electrical Coordination, Canadian Electrical Code, Part III, Canadian Standards Association (CSA), C22.3 No. 3. 50. Questions and Answers About Electric and Magnetic Fields Associated with the Use of Electric Power, National Institute of Environmental Health Studies (NIEHS) and U.S. Department of Energy (DOE), DOE/EE-0040, U.S. Government Printing Office, Washington, DC, January, 1995. 51. School Site Selection and Approval Guide, California State Department of Education, Sacramento, CA, 1989. 52. Background on Electromagnetic Fields and Human Health, Electric Power Research Institute, February, 1987. 53. IEEE Standard Procedures for Measurement of Power Frequency Electric and Magnetic Fields From AC Power Lines, IEEE Standard 644-1994, IEEE Power Engineering Society, March 7, 1995. 54. IEEE Guide for the Measurement of Quasi-Static Magnetic and Electric Fields, IEEE Standard 1460-1996, IEEE Standards Coordinating Committee on Non-Ionizing Radiation (SCC28), March 28, 1997. 55. Guidelines for Limiting Exposure to Time-Varying Electric, Magnetic, and Electromagnetic Fields (Up To 300 GHz), International Commission on Non-Ionizing Radiation Protection (ICNIRP), Health Physics, 74: 494-522, 1998. 56. Threshold Limit Values for Chemical Substances and Physical Agents, American Conference of Governmental Industrial Hygienists (ACGIH), Cincinnati, ISBN 1-88-241723-2, 1998. 57. Guidelines for Limiting Exposure to Time-Varying Electric, Magnetic, and Electromagnetic Fields (Up To 300 GHz), International Commission on Non-Ionizing Radiation Protection (ICNIRP), Health Physics, 74: 494-522, 1998. 58. Threshold Limit Values for Chemical Substances and Physical Agents, American Conference of Governmental Industrial Hygienists (ACGIH), Cincinnati, ISBN 1-88-241723-2, 1998. 59. A Summary of Standards for Human Exposure to Electric and Magnetic Fields at Power Frequencies, B.J. Maddock, Joint Working Group 36.01/06, CIGRE Electra No. 179: 51 65, August, 1998. 60. Interim Guidelines on Limits of Exposure to 50/60 Hz Electric and Magnetic Fields (1989), National Health and Medical Research Council, Canberra, 1989. 61. Low-Frequency Electric and Magnetic Fields Permissible Limits of Exposure for the Protection of Persons in the Frequency range 0 Hz to 30 kHz, Austrian Standard S1119, 1994 (in German). 62. Protection Against the Influence of Electrical Fields in the Closeness of Electrical Transmission Systems for 750 kV and Above, CSN 33 2040, Prague, 1979 (in Czech).

7-33

Personnel Safety Considerations

63. Regulations for Safety and Health Protection in the Workplace for Exposure to Electric, Magnetic, or Electromagnetic Fields, Berufsgenossenschaft der Feinmechanik und Electrotechnik, June 1995 (in German). 64. Maximum Limits of Exposure to Electric and Magnetic Fields Generated at the Rated Power Frequency (50 Hz) in Indoor and Outdoor Environments, Decree of the Prime Minister, Gazzetta Ufficiale della Repubblica Italiana, N.104, 1992 (in Italian). 65. Order of the Council of Ministers dates 5 November 1980 in the Matter of Detailed Principles of Protection Against Non-Ionizing Electromagnetic Radiation Harmful to People and the Environment, Law Gazette, no. 25, item 101, pp 277-278, 17 November 1980 (in Polish). 66. Effects of Electromagnetic Fields Caused by EHV Overhead Power Lines, J. Arciszewski and A. Pilatowicz, CIGRE SC36, Warsaw, 1989. 67. Biological Effects of Electromagnetic Fields, Part 2, Frequency Range 10 Hz to 100 kHz, Report of a Working Group, Environment Text Series No. 214, Swiss Federal Office for environment, Forests and Countryside (BUWAL), Berne, 1993 (in German). 68. Limit Values in the Workplace, Schweizerische Unfallversicherungsanstalt (SUVA), Luzern, 1994 (in German). 69. Board Statement on Restrictions on Human Exposure to Static and Time Varying Electromagnetic Fields and Radiation, Documents of the NRPB, 4, 1-69, 1993. 70. Electric Fields of Industrial Frequency, USSR Official State Standard, GOST 12.1.002-84, Moscow, 1984 (in Russian). 71. Human Exposure to Electromagnetic Fields Low Frequency (0 Hz to 10 kHz), European Prestandard ENV 50166-1, CENELEC, Brussels, 1995. 72. Proposal for a Council Directive on the Minimum Health and Safety Requirements Regarding the Exposure of Workers to the Risks Arising from Physical Agents, OJ No. C 230, 3-29, 19.8.94. 73. Interim Guidelines on Limits of Exposure to 50/60 Hz Electric and Magnetic Fields, International Non-Ionizing Radiation Committee of the International Radiation Protection Association, Health Physics, 58, 113-122, 1990. 74. Order of the Ministry of Health, 23 December 1994 (in Polish). 75. Environmental Health Criteria 69: Magnetic fields, World Health Organization, Geneva, 1987.

7-34

8
FIELD MEASUREMENTS
This chapter provides test procedures and other resources to facilitate testing and evaluation of potential ac interference cases. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE.

8-1

Field Measurements

Introduction
This chapter contains a compendium of test procedures associated with the Investigation and Diagnosis chapter and the Diagnostic Flowchart. Each test procedure is designed to produce one or more results or discriminants, each of which acts as an input to the flowchart. Some of these results will be probabilistic, and can only suggest the correct course of action, as there is often too much individual variability in railroad signal equipment installations to produce more definitive guidance with such simple tests. Each test procedure contains an overview, a required equipment list, a list of test objectives, a test procedure, and instructions on interpreting the results of the test. Also included are several reference figures and tables which the reader may find helpful.

General Precautions for all Test Procedures


1. If steady-state rail-to-ground voltages are in excess of 50 volts, or if railroad equipment has been catastrophically damaged in the past, possibly due to power line faults, then insulated electrical gloves should be used when taking these measurements. If there has not been any history of catastrophic equipment damage and the rail-to-rail voltage is determined to be less than 50 volts steady state, then special work rules, such as high voltage rubber gloves are probably not necessary. 2. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 3. Obtain working time on the tracks, if necessary.

8-2

Field Measurements

Tests
Table 8-1 Diagnostic Flow Chart Tests Test Number 1 2 3 4 Description Excessive common mode voltage Measure dominant frequency Check for excessive track circuit unbalance Rail-to-rail ac interference spectrum Page 8-7 8-10 8-13 8-16

Table 8-2 Resolving Track Circuit Unbalance Test Number 5 6 7 8 9 10 Description Coarse Rail Balance Test Rail-Rail and Rail-Ground Spatial Voltage Distribution Test Insulated Joint Test Direct Measurement of Insulated Joint Resistance Alternate Insulated Joint Test No special equipment Arrester/Equalizer Test Page 8-20 8-24 8-28 8-34 8-38 8-40

Table 8-3 Locating Conducted Sources Test Number 11 12 Hardwire Shunt Testing Local Ground/Power Company Neutral Isolation Test Description Page 8-43 8-45

8-3

Field Measurements

General Information
Table 8-4 Voltage versus dBV Equivalence

8-4

Field Measurements

Figure 8-1 Magnetically Induced Rail Voltage

8-5

Field Measurements

Figure 8-2 Effects of Track Circuit Unbalance on Magnetically Induced Voltage

8-6

Field Measurements

Test #1 Excessive Common Mode Voltage Objectives To determine if continuous ac interference voltage at any location exceeds 50 V rms. Such a condition constitutes excessive common-mode voltage, and may constitute a personnel hazard. Overview Excessive common-mode ac interference voltage on railroads usually comes from one of two sources: either a direct conductive path from a voltage source, or magnetic induction from parallel power lines. Most measurements will be rail-to-ground (common mode). But, if the path is magnetic induction, the highest voltages will probably be found across insulated joints. Equipment Required 1. 1 ea. Digital Voltmeter with high-impedance inputs, such as a (Fluke model 87 D.M.M. or equivalent) 2. Alligator test clips and wires (as required). 3. 1 ea. small temporary ground rod or really large screwdriver. Note: An inexpensive ground rod can be obtained at almost any hardware, electrical supply, or Radio Shack store (part #15530, a 4-foot copper-plated ground rod, is available for about $10). 4. Hammer to drive temporary ground rod. Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the general area in which ac interference is suspected. If possible, these measurements should be made near the point where the most-severely affected signal equipment is attached to the rails of the track. 4. Choose a test location on the tracks where you can safely make several voltage measurements while watching for train traffic.

8-7

Field Measurements

5. At the case or signal housing location, determine which track circuit terminals apply to each rail and each side of insulating joints, and find the ground reference terminal on the surge panel. Alternatively, you may drive a temporary ground rod at any location desired, provided that there is adequate soil conductivity. 6. Attach an alligator clip lead to the ground reference (or ground rod or screwdriver) to provide easy access to the ground connection. 7. Set your voltmeter for ac Volts. DO NOT USE the ac + dc Volts setting (if your meter has one), as this can provide misleading results, particularly for the rail-to-rail voltage measurement. Note: Some experts in the field recommend not performing this test at the readily available terminals inside of the equipment enclosure (bungalow, signal case, etc.) because you cannot always guarantee the quality and purity of the ground connection inside an equipment location. A ground connection with foreign voltages and currents on it may significantly affect the readings taken, and may make them more difficult to interpret. However, the available ground terminal at most signal equipment locations is usually better than any temporary ground rod. Temporary connections to exposed rail are also often poor, so using the permanently connected track leads may be preferable. The signal bungalow or case is also a safer location than being out on the track itself, and may eliminate the need for track and time. Recording equipment can also be left in the equipment enclosure for longer term monitoring. 8. Measure the average ac voltages from each rail to ground (at this location) and record them. 9. Measure the ac voltages from one rail end to the other rail end across each insulated joint, and record them. 10. Remove the temporary ground rod (if installed) and remove all other test equipment from the track area. Move a safe distance away from the tracks. Interpretation of Measurements
First Discriminant: Rail-to-Ground Voltage Magnitude

If continuous ac rms voltage at any location exceeds 50 volts, a personnel hazard may exist.
Second Discriminant: Rail-to-Ground Voltage Magnitude

Most cases of true induction on railroads result in rail-to-ground voltages of 50 volts or less, with 10 volts or less being typical. If the magnitude of the rail-to-ground voltages exceeds 10 volts ac rms, the possibility of a conductive connection to a source of ac power should be more seriously considered than if these voltages are below 10 volts.

8-8

Field Measurements

If the source of the problem appears to be conduction, look at every possible connection point from the rail with the highest rail-to-ground voltage to other circuits. Consider all other circuits as potential sources of interference until proven otherwise. The insulation on electric switch heaters (calrods) should be examined especially closely, as they are normally supplied from a higher-voltage (240 V ac) service than most signal equipment, and can apply dangerously high voltages to the rails.

8-9

Field Measurements

Test #2 Measure Dominant Frequency Objectives To determine the dominant frequency of the ac interference. Overview The dominant frequency of ac interference is usually the fundamental (60 Hz or 50 Hz) or the third harmonic (180 Hz or 150 Hz). If the fundamental frequency of the power system is dominant, then it is probable that the source is associated with nearby transmission facilities (higher voltage). If the third harmonic (or another harmonic) is dominant, then it is probable that the source is associated with distribution facilities. Equipment Required 1. 1 ea. Digital Voltmeter with high-impedance inputs, such as a (Fluke 87 D.M.M. or equivalent), or Audio-Frequency Spectrum Analyzer (or Frequency-Selective Voltmeter) 2. Alligator test clips and wires (as required). 3. 1 ea. small temporary ground rod or large screwdriver. Note: An inexpensive ground rod can be obtained at almost any hardware, electrical supply, or Radio Shack store (part #15-530, a 4-foot copper-plated ground rod, is available for about $10). 4. Hammer to drive temporary ground rod. Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the general area in which ac interference is suspected. If possible, these measurements should be made near the point where the most-severely affected signal equipment is attached to the rails of the track. 4. Choose a test location on the tracks where you can safely make several voltage measurements while watching for train traffic. 5. At the case or signal housing location, determine which track circuit terminals apply to each rail and each side of insulating joints, and find the ground reference terminal on the surge panel. Alternatively, you may drive a temporary ground rod at any location desired, provided that there is adequate soil conductivity. 8-10

Field Measurements

6. Attach an alligator clip lead to the ground reference (or ground rod or screwdriver) to provide easy access to the ground connection. 7. Set your voltmeter for ac Volts. DO NOT USE the ac + dc Volts setting (if your meter has one), as this can provide misleading results, particularly for the rail-to-rail voltage measurement. Note: Some experts in the field recommend not performing this test at the readily available terminals inside of the equipment enclosure (bungalow, signal case, etc.) because you cannot always guarantee the quality and purity of the ground connection inside an equipment location. A ground connection with foreign voltages and currents on it may significantly affect the readings taken, and may make them more difficult to interpret. However, the available ground terminal at most signal equipment locations is usually better than any temporary ground rod. Temporary connections to exposed rail are also often poor, so using the permanently connected track leads may be preferable. The signal bungalow or case is also a safer location than being out on the track itself, and may eliminate the need for track and time. Recording equipment can also be left in the equipment enclosure for longer term monitoring. 8. Measure the average ac voltages from each rail to ground (at this location) and record them. 9. Select the frequency mode on the D.M.M. - or, measure the audio frequency (0 to 1000 Hz) spectrum with a spectrum analyzer or frequency-selective voltmeter - and record the results. 10. Remove the temporary ground rod (if installed) and remove all other test equipment from the track area. Move a safe distance away from the tracks.

Figure 8-3 Hz Button on Fluke 87 to Display Dominant Frequency of AC Voltage

8-11

Field Measurements

Figure 8-4 Hand Held Audio Frequency Spectrum Analyzer

Interpretation of Measurements
First Discriminant: Rail-to-Ground Voltage Magnitude

If continuous ac rms voltage at any location exceeds 50 volts, a personnel hazard may exist.
Second Discriminant: Dominant Frequency

The dominant frequency of ac interference is usually the fundamental (60 Hz or 50 Hz) or the third harmonic (180 Hz or 150 Hz). If the level of the fundamental is greater than the level of the third harmonic, then the frequency counter in the D.M.M. will show either 50 of 60 Hz, and it is probable that the source is associated with nearby transmission facilities (higher voltage). If the third harmonic (or another harmonic) is dominant, then it is probable that the source is associated with distribution facilities.

8-12

Field Measurements

Test #3 Check for Excessive Track Circuit Unbalance Objectives To determine if excessive track circuit unbalance is contributing to the problem by converting common mode interference (rail-to-ground) into differential mode interference (rail-to-rail). Overview It is rail-to-rail ac interference that usually affects the operation of railroad signal equipment. Because the coupling mechanisms that allow ac interference into track circuits tend to introduce almost the same levels into both rails, there is usually very little rail-to-rail interference. But, any unbalance in the track circuit will change some of the rail-to-ground voltage into rail-to-rail voltage. For this reason, track circuit balance is very important in areas with higher levels of environmental ac interference. This test determines the percent (%) unbalance of the track circuit. Equipment Required 1. 1 ea. Digital Voltmeter with high-impedance inputs, such as a (Fluke 87 D.M.M. or equivalent) 2. Alligator test clips and wires (as required). 3. 1 ea. small temporary ground rod or really large screwdriver. Note: An inexpensive ground rod can be obtained at almost any hardware, electrical supply, or Radio Shack store (part #15-530, a 4-foot copper-plated ground rod, is available for about $10). 4. Hammer to drive temporary ground rod. Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the general area in which ac interference is suspected. If possible, these measurements should be made near the point where the most-severely affected signal equipment is attached to the rails of the track. 4. If you are near (within 1000 of) a point on the tracks where ac electrical signals are being applied to the track (AFO transmitter, Motion Sensor or Crossing Predictor equipped with an audio-frequency island circuit, coded ac track circuit), these measurements may be easier to make and interpret if all nearby track circuit transmitters are turned off. AFO transmitters

8-13

Field Measurements

and coded track circuits are of particular concern due to their powerful and/or pulse-mode outputs. Or, use a frequency-selective voltmeter tuned to the power fundamental or harmonic frequency, or use a spectrum analyzer. Track circuits can then remain in service. 5. Choose a test location on the tracks where you can safely make several voltage measurements while watching for train traffic. 6. At the case or signal housing location, determine which track circuit terminals apply to each rail and each side of insulating joints, and find the ground reference terminal on the surge panel. Alternatively, you may drive a temporary ground rod at any location desired, provided that there is adequate soil conductivity. 7. Attach an alligator clip lead to the ground reference (or ground rod or screwdriver) to provide easy access to the ground connection. 8. Set your voltmeter for ac Volts. DO NOT USE the ac + dc Volts setting (if your meter has one), as this can provide misleading results, particularly for the rail-to-rail voltage measurement. Note: Some experts in the field recommend not performing this test at the readily available terminals inside of the equipment enclosure (bungalow, signal case, etc.) because you cannot always guarantee the quality and purity of the ground connection inside an equipment location. A ground connection with foreign voltages and currents on it may significantly affect the readings taken, and may make them more difficult to interpret. However, the available ground terminal at most signal equipment locations is usually better than any temporary ground rod. Temporary connections to exposed rail are also often poor, so using the permanently connected track leads may be preferable. The signal bungalow or case is also a safer location than being out on the track itself, and may eliminate the need for track and time. Recording equipment can also be left in the equipment enclosure for longer term monitoring. Note: the following measurements should be taken as close together (in time) as possible in order to minimize error introduced by fluctuating conditions. All three measurements can easily be completed in under a minute. 9. Measure the average ac voltages from each rail to ground (at this location) and record them. 10. Measure the ac voltage rail-to-rail, and record it. 11. Remove the temporary ground rod (if installed) and all other test equipment from the track area. Move a safe distance away from the tracks. 12. The percent unbalance is equal to the rail-to-rail voltage divided by the rail-to-ground voltage, multiplied by 100. (% Unbalance = 100 * Vr-r / Vr-g)

8-14

Field Measurements

Interpretation of Measurements
First Discriminant: Degree of Track Circuit Unbalance

Track circuit unbalance should be less than 10%. If track circuit unbalance is greater than 10%, then the sources of unbalance must be identified and the degree of unbalance reduced. If track circuit unbalance is less than 10%, this is still no guarantee that rail-to-rail ac interference will not cause operational problems. The 1977 AAR/EEI Bluebook specifies a 50V point-to-point maximum (60V really). This is based on OSHA exposure levels. If the 50V is railto-ground, then 10% unbalance results in 5V rail-to-rail. The AREMA signal manual requires that grade crossing electronics withstand 5V or 10V at 60Hz and/or 180Hz from rail to rail. Yet in reality, such equipment is often susceptible to mis-operation with levels of 180Hz as low as 0.1 to 0.2 volts.

8-15

Field Measurements

Test #4 - Spectral Analysis Test Objective This test is designed to yield insight into the nature of the interference, including its probable source and mechanism of coupling and whether the ac interference originates from transmission or distribution sources. It may also rule out ac interference as the source of the problem. Overview Obtaining spectral information from the track can provide enormous insight into the possible existence, specific type, and possible source of ac interference present on railroad tracks. Equipment Required 1. Audio-Frequency Spectrum Analyzer (or Frequency-Selective Voltmeter) 2. Test Leads (as required) 3. AC Power Isolation Transformer (if needed) Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the general area in which ac interference is suspected. If possible, these measurements should be made near the point where the most-severely affected signal equipment is attached to the rails of the track. This measurement may be made at the signal equipment enclosure, and may be made with the Motion Sensor or Crossing Predictor equipment still in operation. Caution: the inputs of most ac-powered audio spectrum analyzers and most frequency-selective voltmeters are connected to the common (chassis or shield) ground of the unit. When ac powered, this is also connected to the ground prong of the power supply cord. Connecting the ground side of an input to a rail will normally place a direct connection to ground on that rail. Under some circumstances, this could lead to wildly inaccurate results, or even test equipment or railroad signal equipment damage if preventative measures are not taken. If possible, use a battery-powered spectrum analyzer.

8-16

Field Measurements

4. Connect the spectrum analyzers input across the two rails of the affected track, being careful not to accidentally ground either rail via the grounded reference terminal of an unbalanced input on the spectrum analyzer (or frequency-selective voltmeter). Use an isolation transformer on the power supply cord of the spectrum analyzer if necessary. 5. Record the spectrum of voltages present between the rails, or at least take specific measurements at the frequencies of interest and record them. The levels of all railroad signal equipment operating frequencies in-use within the section of track should be recorded. The levels of all power-line related frequencies (60 Hz, 120 Hz, 180 Hz, etc.) should also be recorded, including the lowest-frequency power-line harmonic that is above the highest railroad signal equipment operating frequency in use. Caution: Rail-to-ground ac interference levels can exceed the allowable input levels of audio frequency spectrum analyzers. Always check rail-to-ground ac voltage with a D.M.M. before connecting spectrum analyzers. Using a 10x oscilloscope probe with an audio frequency spectrum analyzer can be an effective precaution. (Recalibration of the spectrum analyzer may be required due to the impedance mismatch between the analyzer and the 10x probe.) 6. Repeat steps 4 and 5 measuring the rail-to-ground spectrum of each rail. 7. Remove the temporary ground rod (if installed) and all other test equipment from the track area. Move a safe distance away from the tracks. Interpretation
First Discriminant: Ratio of AC Power Fundamental to First Odd Harmonic

If the strength of the 180 Hz component (third harmonic) is near or higher than the 60 Hz fundamental, there is a strong likelihood that the source of the ac interference is a nearby power distribution line, and not a transmission line (see Figure 8-5). If the strength of the 180 Hz component (third harmonic) is substantially below the level of the 60 Hz fundamental, there is a strong likelihood that the source of the ac interference is a nearby transmission line, or is from a distribution line with extremely well behaved loads. Note: It is possible that ac interference from both transmission and distribution sources contribute in approximately equal proportions. If uncertain, look at the spectrum of nearby power lines. The distribution spectrum can be measured using a clamp-on current probe with the spectrum analyzer on a pole ground wire. The spectrum of a transmission line can best be measured using a spectrum analyzer connected to a magnetic field coil held directly under the line. If the rail-to-ground spectrum from the track matches one of these two spectra, then you know which is the source. If the rail spectrum has harmonic levels (relative to the fundamental) between these two extremes, then both sources may be contributing to the problem.

8-17

Field Measurements

Figure 8-5 Spectrum of Rail-to-Ground Voltage Caused by AC Distribution Line

Second Discriminant: Fundamental to First Even Harmonic Ratio

If the strength of the 120 Hz second harmonic component (first even harmonic) is not at least 10 dB below the level of the 60 Hz fundamental, there is a strong likelihood that the source of the ac interference is a noisy rectifier of some kind connected to the track1. This can be verified by turning off each rectifier in turn, while monitoring the level of 120 Hz on the track. Once the suspected rectifier (power supply) has been identified, it should be tested by checking its output for excessive ac ripple while the rectifier is connected to a resistive load that draws the same amount of current as the track. If defective, the rectifier must be repaired or replaced.
Third Discriminant: Power and Proximity to Motion Sensor or Crossing Predictor Operating Frequency

The presence of an ac interference rail-to-rail voltage can only affect the operation of track circuits such as a Motion Sensor or Crossing Predictor when the interfering signal has sufficient power and/or proximity (in frequency) to the on-track operating frequency of the railroad signal equipment. This is generally a function of the frequency selectivity of the track circuits receiver. However, if it has sufficient power, ac interference energy can affect track circuits operating at almost any frequency. The specifications vary by equipment manufacturer and specific application.

There is one style of track circuit, known as an AC-DC or Style C track circuit that applies an ac voltage to the rails, and has a diode connected across the track some distance away. This is not the rectifier being referred to here. The rectifier in this case is part of the power supply used for a conventional dc track circuit.

8-18

Field Measurements

A rough rule of thumb for Motion Sensors and Crossing Predictors is to look for ac interference frequencies (the power-line fundamental or its harmonics) that are within 25% of the operating frequency of a Motion Sensor or Crossing Predictor. Within this frequency range, interfering signals stronger than 10 dB below the level of the operating frequency of the Motion Sensor or Crossing Predictor have a potential for causing mis-operation. If the interference energy is lower in power than this, it will generally not interfere with the operation of these track circuits. If the interfering voltage on the track is roughly the same magnitude (i.e., roughly the same dB level as signal equipment operating frequency), abnormal operations are likely, with mis-operation becoming almost assured when the level of interference is 10 dB or more above the level of the signal equipment operating frequency. If the interference energy is outside of 25% of the operating frequency of a motion sensor or crossing predictor, it generally will not cause operational problems until the interference level is high enough to saturate the track circuits track transformers.

8-19

Field Measurements

Test #5 Coarse Rail Balance Test Objectives To make a rough determination of the amount of voltage present on the rails near the affected signal equipment, which may either indicate the presence of a conductive path between a rail and the ac power grid, or the presence of a track unbalance which is converting induced common-mode ac interference voltages into differential-mode (rail-to-rail) voltages that can interfere with the operation of railroad signal equipment. Overview Railroad rails can be unbalanced in one of two ways. First, they can be unequal in their degree of connection to ground (remote earth potential). Secondly, they can be unequal in their degree of connection to a source of ac voltage that is referenced to ground. The first case will produce unequal induced voltages in the presence of an electromagnetic field. The second will produce unequal voltages without the presence of an external electromagnetic field. This test can help determine which of these two processes is occurring, if any. Equipment Required 1. 1 ea. Digital Voltmeter with high-impedance inputs, such as a (Fluke 87 D.M.M. or equivalent) 2. Alligator test clips and wires (as required). 3. 1 ea. small temporary ground rod or really large screwdriver. Note: An inexpensive ground rod can be obtained at almost any hardware, electrical supply, or Radio Shack store (part #15530, a 4-foot copper-plated ground rod, is available for about $10.) 4. Hammer to drive temporary ground rod. Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the general area in which ac interference is suspected. If possible, these measurements should be made near the point where the most-severely affected signal equipment is attached to the rails of the track.

8-20

Field Measurements

4. If you are near (within 1000 of) a point on the tracks where ac electrical signals are being applied to the track (AFO transmitter, Motion Sensor or Crossing Predictor equipped with an audio-frequency island circuit, coded ac track circuit), these measurements may be easier to make and interpret if all nearby track circuit transmitters are turned off. AFO transmitters and coded track circuits are of particular concern due to their powerful and/or pulse-mode outputs. Or, use a frequency-selective voltmeter tuned to the power fundamental or harmonic frequency, or use a spectrum analyzer. Track circuits can then remain in service. 5. Choose a test location on the tracks where you can safely make several voltage measurements while watching for train traffic. 6. At the case or signal housing location, determine which track circuit terminals apply to each rail and each side of insulating joints, and find the ground reference terminal on the surge panel. Alternatively, you may drive a temporary ground rod at any location desired, provided that there is adequate soil conductivity. 7. Attach an alligator clip lead to the ground reference (or ground rod or screwdriver) to provide easy access to the ground connection. 8. Set your voltmeter for ac Volts. DO NOT USE the ac + dc Volts setting (if your meter has one), as this can provide misleading results, particularly for the rail-to-rail voltage measurement. Note: the following measurements should be taken as close together (in time) as possible in order to minimize error introduced by fluctuating conditions. All three measurements can easily be completed in under a minute. Note: Some experts in the field recommend not performing this test at the readily available terminals inside of the equipment enclosure (bungalow, signal case, etc.) because you cannot always guarantee the quality and purity of the ground connection inside an equipment location. A ground connection with foreign voltages and currents on it may significantly affect the readings taken, and may make them more difficult to interpret. However, the available ground terminal at most signal equipment locations is usually better than any temporary ground rod. Temporary connections to exposed rail are also often poor, so using the permanently connected track leads may be preferable. The signal bungalow or case is also a safer location than being out on the track itself, and may eliminate the need for track and time. Recording equipment can also be left in the equipment enclosure for longer term monitoring. 9. Measure the average ac voltage from one rail to the other and record it. 10. Measure the average ac voltages from each rail to ground (at the same location as the previous measurement) and record them. 11. Remove the temporary ground rod (if installed) and all other test equipment from the track area. Move a safe distance away from the tracks.

8-21

Field Measurements

Interpretation of Measurements
First Discriminant: Rail-to-Rail Voltage Magnitude

If the ac voltage from rail-to-rail is less than 0.4 volts rms during periods of abnormal railroad signal equipment operation, this strongly suggests that ac interference is probably not the cause of the mis-operation.
Second Discriminant: Rail-to-Ground Voltage Magnitude

If the ac voltage on each rail, with respect to ground, is less than 10 volts rms during periods of abnormal railroad signal equipment operation, this strongly suggests that ac interference is not the root cause of the mis-operation.
Third Discriminant: Rail-to-Ground Voltage Magnitude

Most cases of true induction on railroads result in rail-to-ground voltages of 50 volts or less, with 10 volts or less being typical. If the magnitude of the rail-to-ground voltages exceeds 10 volts ac rms, the possibility of a conductive connection to a source of ac power should be more seriously considered than if these voltages are below 10 volts.
Fourth Discriminant: Rail-to-Ground Voltage Balance

Severely unbalanced track will be characterized by ac rail-to-ground voltages that are grossly unequal. If the ac interference voltage is truly a result of electric or magnetic induction from nearby power lines, both rails would normally be affected equally, and should show approximately the same rail-to-ground potential, roughly within 10% of one another. The presence of an unbalance between the ac rail-to-ground voltages of approximately 30% or more is an indication that the track is unbalanced in some way, and one of the following two scenarios is likely occurring: a. The unbalance in voltages is the result of an unintended connection from one rail to a source of ac energy such as a hot neutral (via a shorted arrester). In this case the source of the unbalance and the source of the energy are the same, i.e. the unintended connection. This is a case of conduction, not induction. b. The rail-to-ground voltages are being caused by induction, but there is an unbalance between the rails in terms of their individual impedance to ground. This can be caused by one or more shorted insulated joints, a shorted rail-to-ground track arrester, or any other connection from rail-to-ground or from rail-to-rail around an insulated joint that affects one rail of the track more than the other. If the source of the problem appears to be conduction, look at every possible connection point between the rail with the highest rail-to-ground voltage and other circuits. Consider all circuits as suspect until proven otherwise. The insulation on electric switch heaters (calrods) should be 8-22

Field Measurements

looked at especially closely, as they are normally supplied from a higher-voltage (240 V ac) service than most signal equipment, and can apply dangerously high voltages to the rails.
Fifth Discriminant: Voltage Summation

In most cases, the sum of the rail-to-rail voltage and the smaller of the two rail-to-ground voltages should roughly equal the larger of the two rail-to-ground voltages. If this is not the case, then one of two things is most likely happening: 1. the quality of the measurements is suspect, and the test should be repeated, 2. there are phase differences between the voltages (not directly measurable with a D.M.M.), and the cause or nature of the unbalance between the two rails most likely has some other reactive component involved (capacitor, inductor, etc.) Each of the discriminants above provides some small piece of information, which will hopefully lead to an accurate diagnosis. Based on these results, we should be able to narrow the focus of our investigation, and select additional tests to verify the initial clues provided by this test.

8-23

Field Measurements

Test #6 Rail-Rail and Rail-Ground Spatial Voltage Distribution Test Objective To make a determination of the distribution pattern of the voltage on the rails near the affected signal equipment, which may indicate the presence and/or location of an unbalancing track anomaly. This may be a conductive path between a rail and the ac power grid, a track unbalance that is converting induced common-mode ac interference voltages into differential-mode voltages that can interfere with the operation of railroad signal equipment. Overview Measuring the ac voltages from rail-to-rail and from rail-to-ground along the length of the affected section of railroad track can provide clues as to the type and location of a failed track circuit component which is creating an unbalance in the track. Equipment Required 1. 1 ea. Digital Voltmeter with high-impedance inputs, such as a (Fluke 87 D.M.M. or equivalent) 2. 1 ea. Small temporary ground rod or really large screwdriver. Note: An inexpensive ground rod can be obtained at almost any hardware, electrical supply, or Radio Shack store (e.g. Radio Shack part #15-530, a 4-foot copper-plated ground rod, works well for this, and is available for about $10 USD.) 3. Hammer to drive temporary ground rod. 4. Alligator test clips and wires (as required). 5. Contractors or surveyors Measuring Wheel (for measuring distances along the rail). 6. Graph Paper/Pencil or Pen (May also plot using computer graphing or spreadsheet software) Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items, such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the general area in which ac interference is suspected, and determine the location of the endpoints of the track circuit involved. Note: the track circuit does not necessarily end at 8-24

Field Measurements

the first set of insulated joints encountered in each direction when moving away from the middle of the circuit. The track circuit ends are defined by the first pair of insulated joints which have been established to be in good working order (no shorts) and which are not bypassed by any device capable of carrying the interfering frequency around the insulated joints. Such devices include all tunable joint couplers, wideband couplers, hardwire couplers, and other accidental means of connection such as multiple failed lightning arresters. 4. If possible, turn off all signal equipment connected to the track within the portion of track to be measured. This is often a nearly impossible task, but it will greatly clarify the meaning of the measurements taken. Or alternatively, use a spectrum analyzer such as in test #4 and measure the power frequency and its harmonics. 5. Begin at one end of the track circuit (as defined above), and drive a temporary ground rod into the soil or ballast nearby. Ideally, this will be driven into moist soil somewhere within a few feet of the tracks, but outside of the ballast. If a suitable temporary ground rod is unavailable, and the soil is adequately moist, the shaft of a really large screwdriver may be shoved into the ground with just enough of the metal protruding above ground to attach one of the voltmeter leads. 6. Measure and record the rail-to-rail voltage, as well as the rail-to-ground voltages for each rail. 7. Remove the temporary ground rod (or screwdriver). 8. Using a contractors measuring wheel, measure off one tenth (or less) of the length of the track circuit. Note: at least eleven measurements should be taken within the track circuit, in order to provide sufficient resolution in the graph that will be plotted at the end of the test. 9. Install the temporary ground rod at the new location, and repeat the rail-rail voltage measurement, as well as both rail-ground measurements. Record the results. 10. When measurements have been taken and recorded along the entire length of the track circuit, plot them on graph paper or a computer. The general shape of the results on a normal section of track with no operating signal equipment should resemble the figure below:

8-25

Field Measurements

Figure 8-6 Magnetically Induced AC Rail Voltages

Interpretation Notice that the sample graph above shows a generally low and somewhat random voltage for the rail-to-rail measurement, with the rail-to-ground measurements having a V shape. The magnitude of the voltages here is arbitrary, and may not reflect the actual magnitudes measured in any given track circuit. It is the general shape of the curve that is important. As turning off all signal equipment within a particular section of track may be extremely difficult to arrange, the data gathered at a particular location will have to be interpreted with an understanding of the test conditions that were provided. Any track circuit or other equipment, which applies ac voltages to the rail, can cause a localized or widely distributed upward (higher voltage) deviation from the idealized figure above. In general, the higher the frequency of the track circuit, the more localized the upwards voltage deviation will be. For a normal section of railroad track, the induced ac interference voltage will generally appear to be highest (rail-to-ground) at both ends of the track, with the lowest rail-to-ground voltage found at or near the middle of the section. (This assumes an essentially constant electromagnetic field.) Of course, if the electromagnetic field is not constant along the affected section of track, the voltage from rail-to-ground will be affected by this.

8-26

Field Measurements

First Discriminant: Rail-to-Ground Voltage

If the ac interference rail-to-ground voltage on one end (at the insulated joints) of one or both rails is significantly (a few volts to several volts) higher than the rail-to-ground voltage at the opposite end, this is an indication that the insulated joint(s) at the low rail-to-ground voltage are shorted or bypassed by something which can conduct the interference voltage around the insulation. An insulated joint test should be performed, beginning with the insulated joint on the rail with the low rail-to-ground voltage. If the insulated joint is found to be defective, it must be repaired or replaced, and Test #2 should be repeated.
Second Discriminant: Rail-to-Rail Voltage

If the rail-to-ground voltage is higher at one end of the section of track than the other, and the rail-to-rail voltage is low throughout, this is an indication that both of the insulated joints at the low rail-to-ground voltage end of the section of track may be shorted, and should be tested (see test #7).
Third Discriminant: Rail-to-Ground Voltage Anomalies

If the Rail-to-Ground voltage is highest at some point other than an end of the section of track, this is an indication of a source of voltage that is not uniformly distributed along the section of track. This can occur when the source of voltage is a conductive path, or when the area of exposure to the electromagnetic field causing the ac interference voltage is less than the entire section of track (non-uniform inductive excition). When there is a mid-section voltage peak that affects both rails approximately equally, this is an indication of induced ac interference energy. The peak in rail-to-ground voltage will be near the center of the electromagnetic field causing the induction. When there is a mid-section voltage peak that affects only one rail, or affects one rail much more than the other, this is an indication of conducted ac interference energy. The rail-to-ground voltage will be highest at the point nearest the source of conducted energy.

8-27

Field Measurements

Test #7 Insulated Joint Test Objective To determine whether an insulated joint has failed (shorted-out or low-resistance), and should be repaired or replaced. Overview Failed insulated joints are frequently a significant source of track unbalance. Many railroads fail to understand the importance of having all insulated joints surrounding an isolated track circuit working properly. Normally, track circuits will not fail under ordinary circumstances unless two or more insulated joints have failed. Railroads often continue to operate their signal systems with many latent insulated joint failures. In the presence of an electromagnetic field, such as that which can be created by nearby ac power lines, these failed insulated joints can cause an unbalancing of the track, which can convert the normal and inconsequential presence of induced common-mode ac voltages into differential-mode voltages which appear as a difference in potential between the two rails of the track. Such differential-mode voltages can directly affect the normal operation of track circuits. The following graph shows how a degraded insulated joint can convert common mode interference into differential mode interference.

Figure 8-7 Degraded IJ Converts Common Mode Interference into Differential Mode

8-28

Field Measurements

When ac power lines and railroad tracks share a corridor, the importance of properly testing and maintaining insulated joints cannot be over-stated. Unfortunately, this process is often difficult to perform accurately with the joint installed in the track. And even though there are often governmental regulations that apply to the inspection of insulated joints, enforcement of these regulations has historically been lax, with no periodic testing requirements. However, the process of testing insulated joints is very important, as having a complete set of properly working insulated joints to define the absolute physical and electrical limits of a track circuit is essential to the avoidance of ac interference problems. Failed (shorted or at least partially-conductive) insulated joints, along with shorted lightning arresters, are the leading hardware failures that can act as the catalysts needed to unbalance a railroad track circuit and produce an ac interference problem in a railroad track circuit. Testing of installed insulated joints cannot be done using a conventional Ohmmeter or insulation tester, as the internal sources used by these instruments are unsuitable. In the case of the insulation tester, the presence of the inevitable ballast conductance will make the insulation appear to have failed, even when it is good. And the voltage of the internal source in an Ohmmeter can be as low as a tenth of a volt for some models. This will not consistently provide enough voltage to forward-bias the metallic short-circuits which can form inside of an insulated joint. These shorts can exhibit a semiconductor characteristic, which is not unlike a diode, in that no current will flow (the insulation appears good) until a voltage potential of several tenths of a volt (or more) has been applied. The degree of failure (conductivity) required for an insulated joint to cause abnormal operation of a track circuit depends on many variables, including: 1. The characteristics of the track circuit itself, including its physical layout. 2. The strength of the electromagnetic field present 3. The length of the track circuits on either side of the failed insulated joint 4. The ballast resistivity in the track circuits 5. The presence of other track circuits or track circuit components connected to the rails. Equipment Required 1. Railroad-specified insulated joint test equipment or test set 2. (Optional) Insulated joint test instrument from a test equipment manufacturer 3. (Optional) Insulated joint tester consisting of a 6-Volt lantern battery, a 10-Ohm, 5-Watt (minimum) resistor, and an ammeter capable of accurately measuring at least one Ampere of current (the current-measuring mode of an analog multimeter or a D.M.M. will do).

8-29

Field Measurements

Procedure Unlike the other test procedures in this chapter, no single, specific test procedure to be followed will be outlined here. Almost every railroad in North America has its own standards and procedures for the testing of insulated joints, and these are to be preferred as the primary means of determining the condition and suitability of an insulated joint in almost every case. As these procedures are not used frequently enough, their existence is sometimes forgotten by signal maintenance personnel in the field. Additional inquiries directed to the railroads signal maintenance personnel and department office may be required in order to make them available. It is not the intent of this handbook to attempt to define an optimum test procedure for testing insulated joints. In all cases, the prevailing railroad test standards, if any, must be considered as the first criteria for evaluating the condition of an insulated joint. However, where these standards are lacking, or do not follow both the intent and the specific requirements of the applicable regulatory agency standards, the applicable regulatory agency standards should be applied (e.g. F.R.A. rules: 234.205, 234.235, 236.59, 236.735, and 236.752 in the United States). The applicable railroad test procedure, if available, should be followed in the testing of insulated joints. Alternatively, a purpose-built insulated joint testing instrument should be used in accordance with the manufacturers instructions. If neither of these options are available, the following procedure may be used in order to perform a basic test of an insulated joint. 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the suspected insulated joint. 4. Connect an Ammeter capable of accurately measuring a current of 1 Ampere, a 10-Ohm/5Watt Resistor, and a 6-Volt Lantern Battery to form a simple insulated joint tester as shown below:

8-30

Field Measurements

Figure 8-8 Insulated Joint Tester

1. Temporarily short the two test leads together and verify that the Ammeter reads approximately 550 milliamperes, or more. Otherwise, replace the defective component or repair the wiring of the circuit. 2. Connect the insulated joint tester constructed in step 1), above, across the insulated joint as shown in the figure below: 3. Record the amount of current flow indicated by the ammeter. 4. Disconnect the test apparatus from the rails on either side of the insulated joint.

8-31

Field Measurements

Figure 8-9 Use of Insulated Joint Tester

Interpretation The results of a standard railroad test for insulated joints should be interpreted using the guidelines provided by the railroad test procedure. Similarly, results obtained using an insulated joint test instrument should be interpreted using the test instrument manufacturers guidelines. The results of the basic insulated joint test procedure described above can be affected by many external factors, but are primarily affected by ballast resistivity. Ballast resistivity can vary over a very wide range, and there is no typical value that can be used in calculating allowable insulated joint leakage currents when measured in this fashion. This considerably restricts the degree of certainty with which we can interpret the results of this test. Therefore the following broad guidelines are suggested as a means of interpreting the results of this test over a wide range of ballast conditions, with no implication that they represent hard and fast rules for testing insulated joints: 1. Any insulated joint, which produces a current reading of greater than or equal to 300 milliamperes, should be considered defective, until proven otherwise. 2. Any insulated joint, which produces a current reading of between 100 and 300 milliamperes, should be considered suspect, until proven otherwise. 3. Any insulated joint, which produces a current reading of between 50 and 100 milliamperes, may either be defective, or may be installed in less-than-ideal ballast resistivity conditions. Without additional information about the resistivity of the ballast, no further conclusions can be drawn from this result.

8-32

Field Measurements

4. Any insulated joint, which produces a current reading of 50 milliamperes or less, is probably good. The greatest exception to this will be in conditions of high ballast resistivity, where the majority of the leakage current will actually be carried by a failed insulated joint, and not by the surrounding ballast. The effectiveness of the insulation in an insulated joint is the only discriminant produced by this test, and even this result is significantly clouded by the effects of ballast resistivity2. It is the personal opinion of the principal author(s) of this chapter that when ac interference problems are encountered, all bad or suspect insulated joints should be replaced, repaired, or at the very least - investigated further. The replacement cost for insulated joints usually pales in comparison to the amount of expense incurred by conducting a lengthy investigation of other causative factors and potential solutions.

Most insulated joint tests are actually testing the resistance of two parallel paths. One path is the insulated joint; the other is from one rail through the ballast/earth to the other rail. If one path resistance is much lower than the other, then the measurement will reflect that lower resistance. Assuming ballast resistance is higher, the current measurement ranges specified above correspond to the following insulated joint resistances: 1. 2. 3. 4. Defective Suspect <10 10 to 50 >110

Need ballast info 50 to 110 Probably good

8-33

Field Measurements

Test #8 Direct Measurement of Insulated Joint Resistance Objective To directly measure the resistance of an insulated joint. Overview Because the presence of an alternate path for current when testing installed insulated joints, it is sometimes worth the effort to make a more direct measurement that eliminates the confounding factor of low ballast resistivity. Such a measurement could be used when more common measurement techniques produce uncertain results. Note: This measurement is only effective if at least some ac energy is present on the rail. Equipment Required 1. 1 ea. Digital Voltmeter with high-impedance inputs, such as a (Fluke 87 D.M.M. or equivalent)(2 D.M.M.s are better) 2. Alligator test clips and wires (as required). 3. 1 ea. Flexible Current Probe, 36, maximum available sensitivity (such as the AEMC AmpFlex 1000-36-2-1, LEM-Flex RR3035/36, or equivalent) Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Place the current probe around the suspect insulated joint between the track leads. 4. Measure the ac current through the insulated joint using the D.M.M. (have the current probe on its most sensitive setting highest mV/A) 5. Measure the ac voltage across the insulated joint using the D.M.M. (steps 4 and 5 can be performed simultaneously if two D.M.M.s are available) 6. Divide the voltage reading (volts) by the current reading (amps) to determine the resistance (ohms).

8-34

Field Measurements

Figure 8-10 Photograph of IJ Leakage Current Test

Interpretation New insulated joints usually measure 20 to 30 Megohms (20,000,000 to 30,000,000). They frequently measure 300 to 1000 in the field because of ballast and contamination. This is adequate for track circuits that operate at 1 to 3 impedances. 10 is usually adequate to effectively isolate individual track circuits. But, in the presence of magnetic induction, higher resistance is needed to avoid unbalances that can result in abnormal operation of equipment such as grade crossing train detection electronics. Insulated joint impedances below 50 can cause 5% track circuit unbalance (see Figure 8-7). IJ impedance below 30 can cause 10% track circuit unbalance. The following graph shows the relationship between voltage across the joint, current through the joint, resistance of the joint, and track circuit unbalance.

8-35

Field Measurements

Figure 8-11 IJ Resistance from Voltage and Current

The accuracy of these reading gets worse with lower current. Current probes such as the example cited above (AmpFlex 1000-36-2-1) have a reading accuracy of 1% over their calibrated range. That calibrated range has a lower limit of 5A. One such current probe was tested and was accurate to 1% at 1.00A. The reading error increased to 100% at a reading of 0.14A (0.07A actual). Good joints tend to indicate current amplitudes around 0.006A. Bad joints tend to indicate current amplitudes of 0.5A or more with a few volts across the joint. The following graph can be used to accurately correlate readings with resistance and potential circuit unbalance for the current probe tested.

8-36

Field Measurements

Figure 8-12 IJ Resistance from Voltage and Current Adjusted for Tested Current Probe

8-37

Field Measurements

Test #9 Alternate Insulated Joint Test No Special Equipment Objective To test for a shorted insulated joint Overview If an insulated joint is suspected of causing rail-to-rail interference by unbalancing the track circuit, and if no special test equipment is available, this quick test can provide an initial indication of a shorted insulated joint. Equipment Required 1. 1 ea. Digital Voltmeter with high-impedance inputs, such as a (Fluke 87 D.M.M. or equivalent). 2. Alligator test clips and wires (as required). 3. A hardwire Test Shunt, in good working condition. Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Measure the ac voltage across each insulated joint of the nearest pair of insulated joints, and at the first set of joints beyond it in each direction. 4. Identify the pair of insulated joints with the greatest difference between their voltages. Note which joint had the highest voltage across it. 5. Monitor the elevated ac voltage from rail-to-rail. 6. Place the hardwire shunt around the insulated joint that had the highest voltage. 7. If the rail-to-rail voltage is effectively eliminated by the presence of the hardwire shunt around the insulated joint, then it is probable that the insulated joint with the lowest voltage (the one without the shunt) is shorted. 8. Remove the shunt. 8-38

Field Measurements

Interpretation Clearly, shorting insulated joints in this way is a brut-force test. But the point of this test is to identify a shorted insulated joint without waiting for a current probe or insulated joint tester to arrive. If a shorted insulated joint is causing the track circuit to be unbalanced, then shorting the accompanying joint should restore balance. The signals wont work while the shunt is in place, but the failed insulated joint will have been identified.

8-39

Field Measurements

Test #10 Arrester/Equalizer Test Objective To determine the status (good/bad) of a lightning arrester or equalizer. Overview Failed (shorted) lightning arresters connected to railroad tracks and other circuits can be a source of unbalance. As lightning arresters in railroad signal circuits are connected from the rail, wire, or circuit path to be protected, to an earth ground (or some other voltage reference), a shorted failure mode in the arrester effectively connects that rail, wire, or circuit path to ground. Different types of lightning arresters have different current-handling capabilities, firing voltages, and failure characteristics. Some types of arresters are not widely accepted for use on railroad track circuits, such as gas-tube arresters, due to their tendency to fail in a shorted fashion. However, the beneficial low-firing-voltage characteristics of gas-tube arresters still tempts many railroads to use them for protecting track circuits. The potential danger here is that the presence of two shorted arresters (one from each rail to ground), could cause a false shunt across the tracks, which would be low enough in resistance to effectively blind a Motion Sensor or Crossing Predictor to the presence of the oncoming train, but would not have a low enough resistance to make the track circuit of the wayside signaling system appear occupied. This could result in a train approaching a grade crossing at more than a restricted (slow) speed, when the crossing warning equipment would be unable to provide adequate warning time. Other protective devices, often known generically as equalizers, are special arresters having a lower firing voltage than arresters. These are normally connected from one rail to the other of the same track. In the event of a nearby lightning strike, or other over-voltage event, the equalizer is intended to equalize the voltage potential of the two rails, so as to minimize the potential for damage to the signal equipment. Arresters and equalizers should essentially appear as an open circuit, at voltages of 10 volts or less. However, they may contain a high-voltage semiconductor material that will show some slight leakage current at these voltages. An ordinary ohmmeter can be used to determine the failure status of a lightning arrester or equalizer. Equipment Required 1. An analog or digital Ohmmeter (Fluke 87 D.M.M., Simpson 260 or TS 111, or equivalent) 2. A clamp-on AC current meter (Fluke 330-series, or equivalent)

8-40

Field Measurements

3. A means of physically removing the lightning arrester or equalizer from its location on the surge panel of a railroad signal equipment installation. This may require the use of a railroad terminal wrench, or may not require any tools at all for arresters equipped with snap-in socket mounts. 4. Test Leads (as required). Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Remove the lightning arrester or equalizer to be tested from its mounted position on the surge panel in the railroad signal equipment enclosure. Note: once the location of a suspected failed arrestor has been narrowed down to a particular equipment installation by the use of test procedure #5 and or #6, the search for a potentially failed arrestor can be further narrowed by the use of a clamp-on ac current meter. Clamp the current meter around the ground lead from each bank of arrestors in the location to determine if there is ac current flowing. Normally there will be no measurable current flowing through any arrestor or arrestor panel ground lead (<1 mA). The presence of even a few milliamps of ac current in the ground lead of an arrestor panel can indicate the presence of a failed arrestor. Once ac current is found in the ground lead of an arrestor panel, the clamp-on current meter may then be used around each arrestor, or the conductors leading to or from it to identify the specific lightning arrestor which has failed. (A clamp-on ac current meter may also be used around equalizer devices which are connected from rail-to-rail, instead of from rail-to-ground, as the current flow in these devices is also normally zero unless they have failed.) Any surge protective device which appears to be passing even a few milliamps of current while in-circuit should be removed for testing as described below. Care should be used when removing arrestors and equalizers from their mountings so as to avoid loosening or damaging their component parts, which could affect the test results obtained. 4. Connect the ohmmeter across the two leads of the arrester, as shown below. For threeterminal arresters, connect the ohmmeter across the center and outer legs, as shown below. 5. Record the resistance(s) measured across the terminals of the arrester.

8-41

Field Measurements

Figure 8-13 Arrester Testing

Interpretation Due to the low-impedance nature of railroad track circuits, almost anything with an impedance of greater than 10 Ohms will be essentially invisible to the track circuitry. However, this is not a suitable threshold for declaring an arrester to be bad. As the normal resistance of a good arrester or equalizer approaches or exceeds 1 Megohm, a more proper threshold lies somewhere closer to the middle of these extremes. In general, any arrester that shows a measured resistance of less than 1,000 Ohms should be immediately discarded and replaced, unless specified otherwise by its manufacturer. The only discriminant produced by this test is a determination of whether or not an arrester or equalizer has failed in a shorted mode. Typically Arresters will show a greater resistance than Equalizers, but either device may read open when measured on most ohmmeters.

8-42

Field Measurements

Test #11 Hardwire Shunt Testing Objective To locate a potential conducted source of ac interference. Overview When other testing has indicated the possible presence of a conducted source of interference, there is a simple test which can be performed with very little test equipment. Furthermore, this test equipment (a hardwire shunt) will be carried by almost all railroad signal maintenance personnel. This test relies on the general electrical characteristics of railroad tracks and conducted interference sources, and attempts to locate the source of the conducted interference within an electrically isolated section of track. It does this by having the tester place a hardwire shunt across the tracks at various locations within the section of track, and then observe the effects of this on the affected railroad signal equipment. Placing a hardwire test shunt across the tracks forces an equalization of rail-to-rail voltage potentials at that point. This also forces an equalization of all ac interference voltages at all points on the opposite side of the hardwire test shunt from the source of ac interference energy. The effectiveness of the hardwire test shunt will increase as it is placed closer to the source, provided that it is still between the source and the receptor (the railroad track circuit receiver). This effect will reach a maximum when it is located at the source of the conducted interference. The effectiveness of the hardwire test shunt decreases as it gets farther from the source of the conducted interference in either direction from the source, with the effectiveness dropping off more sharply after passing the receptor while proceeding away from the source. This test is primarily used in cases of suspected conducted interference, as the interpretation of results when shunting a track that is being influenced by inducted ac interference is much more complex. Equipment Required 1. A hardwire Test Shunt, in good working condition. 2. A means of monitoring either the level of the ac interference voltage across the tracks, or its effect on the railroad signal equipment. 3. A means of remotely communicating with an assistant (Radios, Cell Phones, etc.)

8-43

Field Measurements

Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the general area in which ac interference is suspected. 4. Start at one end of the affected section of track, which will usually be at a set of insulated joints beyond which there have been no anomalies observed. 5. Apply the hardwire shunt to the track and note the effect of this on the ac interference voltage level, or on the operation of the signal equipment, at the affected location. 6. Remove the hardwire shunt from the track. 7. Repeat steps 5) and 6) at various locations throughout the affected section of track. 8. When the opposite end of the affected section of track has been reached, the test is complete. Interpretation Compare the effectiveness of each hardwire test shunt location in reducing the level or effect of the ac interference at the affected location. If there is a source of conducted interference (implied by other testing such as Test #5, above), then the source of the ac interference will likely be close to the point of maximum hardwire test shunt effectiveness. This test may be repeated, if necessary, in order to zero-in on the exact location of the source. If the source is distributed over a wide area, the results of this test will probably be inconclusive.

8-44

Field Measurements

Test #12 Local Ground/Power Company Neutral Isolation Test Objective To determine whether interference currents may be entering one of the rails from the ac power service Neutral wire. Overview A simple method for detecting the presence of a common type of conducted ac interference is to isolate the Neutral bus of the ac power service from the signal ground bus at the signal equipment installation. The power service Neutral is usually grounded at the service drop pole, and at the breaker box. The signal ground bus is connected to a locally driven ground rod network. Normally, these two buses are tied together at a railroad signal equipment installation. However, if the Neutral bus is at some voltage potential other than ground (local earth or remote earth), then current will flow from the Neutral bus into the signal ground bus. This can energize the signal ground bus, which can often have an excessively high resistance to ground. If there is a shorted arrester from one rail of the track to the signal ground bus, this can energize one rail of the track to an elevated ac interference potential. Often it is easier to remove the single wire connecting the Neutral bus in the circuit breaker box to the signal ground bus, than it is to remove, test, and re-install every track circuit arrester connected to the affected rails. Equipment Required 1. Miscellaneous hand tools, as required, to disconnect the wire from the Neutral bus in the circuit breaker box to the signal ground bus. 2. A means of monitoring either the level of the ac interference voltage across the tracks, or its effect on the railroad signal equipment. Procedure 1. Take all actions necessary to protect your own safety as well as the safe movement of trains, highway traffic, and pedestrians. This includes personal protective equipment, job safety briefings, and any other safety-related items such as arranging for assistance from other personnel, if necessary. 2. Obtain working time on the tracks, if necessary. 3. Locate the signal equipment installation at which ac interference is suspected.

8-45

Field Measurements

4. Wearing high voltage gloves, remove the wire connecting the Neutral bus of the power service to the signal ground bus. Remember that the neutral wire carries some current and may be at an elevated voltage, so caution must be taken to ensure that the disconnected neutral does not make contact with other equipment. 5. Observe either the operation of the signal equipment, or the level(s) of the ac interference voltage on the rails of the track. Interpretation If removing the wire between the Neutral bus of the power service and the signal ground bus reduces the level of ac interference voltage on either or both of the rails of the track, this is an indication that there is an unintended connection between the signal ground bus and one or more of the rails. Usually this will be in the form of a shorted lightning arrester from rail to ground. A thorough testing of the lightning arresters should be conducted to find the failed arrester(s), if any. If there are no shorted lightning arresters, then the presence of another path from the power company neutral to one rail must be assumed. Minimizing or eliminating this conductive path will be necessary in order to minimize the voltage on the rails.

8-46

9
INVESTIGATION
This chapter provides an effective philosophy, based on the scientific method, for investigating potential ac interference conditions on railroads. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE.

9-1

Investigation

Introduction
The process of investigating ac interference problems on railroads must be an organized study designed to achieve a complete understanding of the sources, paths, and receptors of the unwanted energy. This organized study needs to be backed by an understanding of the principles of electromagnetic compatibility (EMC), the systems that railroads employ to monitor and control trains and vehicular traffic, and the relevant aspects of the electric power system. Building upon these basics, it is important to develop experience in recognizing related events that can indicate the presence of EMI problems. Chapters 1 through 4 of this handbook provide some of the basic understanding that is needed. Chapters 8 through 12 provide the basis of the organized study. For problems involving existing installations, this process is based on the diagnostic flow chart of Chapter 10. For evaluations of proposed installations, Chapters 11 and 12 provide the process. And, Chapters 5 through 7, 13, and 14 provide shared experience on which the reader can build. In this chapter we present the organized thought processes necessary to weave these strands together.

Quick-Start Version
The quick-start version of this investigation chapter should be used with great care, as it is really just a quick-reference outline for experienced personnel. Troubleshooting Fundamentals The Scientific Method The process of troubleshooting is essentially an applied version of the scientific method: 1. Recognize or identify a problem (or question) 2. Study the available information about the problem or question 3. Guess at a possible cause (or answer) 4. Predict the consequences and implications of that cause or answer 5. Devise and perform an experiment to test those predictions 6. Form the simplest possible theory that explains your guess, your prediction(s), and the experimental results. Step 1): Recognition A Report of a Problem This is when we put on our detective hats and start asking the basic who/what/where/when questions like: 1. Who reported this? 2. What happened? 9-2

Investigation

3. Where did it happen? 4. When did it happen? 5. Is it still happening? 6. When did it stop? 7. In what way was this different than the expected behavior of the system? 8. What should have been happening at that time? 9. Has this ever happened before? 10. Have there been any already-identified causes of this unusual behavior? 11. Is there anything unusual taking place that may have caused this? 12. What has changed on the railway or power system? But the first major decision made during the investigation must always be: Did something really happen? Step 2): Study the Problem At this point we are not concerned with the causes, only the symptoms, i.e. the precise nature of the unusual behavior of the system. We simply need to begin with a clear definition of the anomaly. a. Question b. Examine c. Decide What type of deviation from normal system behavior was observed? Did we review trouble reports of the railway and power company for this area and what information can we glean from them? (We can take some preliminary rail-to-rail and rail-to-ground voltage measurements as described in the first part of Chapter 8 and Chapter 10.) Step 3): Guess at the Cause(s) When we decide that we have enough details about the problem to make an intelligent guess (called a hypothesis) at the cause(s) of the problem we do so. Sometimes we may come up with several competing hypotheses simultaneously. Each of which must be tested. This is often a key point at which serious mistakes are made. Usually, the mistakes result from not considering enough possible causes of the symptoms. 9-3

Investigation

Step 4): Predict the Implications Now that we have one or more guesses or hypotheses, we need to refine one or more of them into a theory. The purpose of a theory is to explain the experimental data that we observe. In order to turn a hypothesis into a theory, we must identify the implications of the hypothesis, and then test them. This is the step in the scientific method that requires us to flesh out our hypotheses in preparation for testing them. Beginning with our preliminary observations, and taking each of our hypotheses in turn, the logical implications of each hypothesis must be identified. These identified implications will form the basis of our tests. Step 5): Test Each Implication We will perform the tests that will most quickly discriminate between the various competing hypotheses, and identify those hypotheses that most correctly explain the data. Step 6): Form the Simplest Theory that Explains the Results Once we have conducted a test, we must then decide what the results mean. If our results confirm our hypothesis, then we would tend to accept it as a working theory, and either go on to look for more ways of confirming or disproving it, or proceed to apply the theory in order to find a solution to our problem. Being sure that you have enough data to confirm your hypothesis and then proceeding to apply your newly developed theory is often difficult. This decision is usually a judgment call hopefully one based on past experience. If our results do not confirm a hypothesis, then we must abandon it, and test another hypothesis. Often, cases of ac interference are correctly diagnosed only after many other hypotheses have been discredited. Unfortunately, the outward symptoms of ac interference can be very similar to those resulting from many other causes of railroad signal trouble. Is it Really AC Interference? We should look for the common and ordinary causes of trouble first, before we resort to less-probable hypotheses. Many signal system problems are too-quickly attributed to ac interference, high-tension power lines, or induction when in fact the trouble has actually resulted from ordinary signal equipment failures or intermittencies. We first need to determine two things: 1. We need to verify that there is no underlying failure in the signal equipment that would have been detected by standard railroad signal maintenance and troubleshooting practices, and that all components of the signaling equipment are in good working order. 2. We need to have positively identified the specific equipment or system that is causing the observed problem.

9-4

Investigation

The Rules of Thumb of Railroad Signals and AC Interference These are the basic facts and prevailing opinions that should guide your investigation of a suspected case of ac interference with railroad signal systems. 1. Ninety percent of everything is track. 2. Motion Sensors and Crossing Predictors are the most sensitive devices connected to the track, and will usually be the first indicators or victims of ac interference. 3. Railroad signal circuits see the rails of the track differentially. 4. Anything which unbalances the electrical characteristics of one rail with respect to the other can act as a catalyst in turning induced common-mode voltage into rail-to-rail voltage. 5. Generally speaking, a lot of cases of induction in reality turn out to be conduction. 6. Just because theres some ac interference on the track, this doesnt mean that ac interference is the cause of the problem. 7. Many railroad signal circuits are frequency-selective, but enough of even a non-adjacent frequency can cause operational problems. 8. When you hear hoof-beats, expect horses, not zebras. 9. The reliability of a railroad signal component may be proportional to its distance from the track. 10. Transmission lines may look big and menacing, but ordinary distribution lines are more often the source of ac interference.

Detailed Version
Introduction The process of investigating and diagnosing suspected cases of ac interference on railroads is one of the least understood aspects of railroad signal maintenance. This simple fact has been the primary motivation for writing this handbook. In this chapter, we will attempt to explain the process of investigating potential ac interference problems involving power facilities and railroad signal equipment. Although ac interference can be an expensive and frustrating headache for both power companies and railroads alike, it is important to keep the actual scope of this problem in perspective. AC interference affecting railroad signal systems is poorly understood by both railroads and power companies alike, because of its relative rarity. Despite the thousands of miles of shared corridor exposure, there are relatively few verifiable instances of significant ac interference presently occurring in North America.

9-5

Investigation

This relatively low rate of occurrence results in very few individuals having had the opportunity to witness and/or be involved with the repair of many such problems. Power company engineers are very familiar with unintended ac induction, but generally do not fully understand its effects on railroad signaling equipment. And railroad signal maintenance personnel normally do not have much exposure to ac power transmission and distribution practice, nor do they fully understand the physics of electromagnetic induction. Thus, the specialists in each area remain most comfortable within their own respective worlds, and the sorely needed cross-pollination rarely occurs. The purpose of this chapter is to develop investigative tools that can be used by railroad and power company personnel to investigate ac interference problems. We will begin with a general discussion of the troubleshooting process, introduce the source-path-receptor model, provide some general guidance, and then discuss a suggested troubleshooting process in greater detail. The Nature of the Problem Once a signal system has been properly designed and installed on a railroad, its operation is generally quite reliable. Maintenance consists primarily of periodic testing and inspection to make sure that no latent failures have occurred (such as unintentional or false grounds), and making sure that the storage batteries (used for battery back-up) are in good condition. Aside from this, the repair or replacement of failed or damaged components on an as-needed basis is usually all that is required. Unfortunately, ac interference problems lie outside the boundaries of this routine. They can occur even when there has been no failure in any of the hardware or software used in the railroad signal system or ac power system. And current federal regulations in the United States and Canada do not require regular testing for the presence of ac interference. These problems can arise from something as simple as a change in the characteristics of the load on the ac power system. They may come and go on their own schedule, with little apparent rhyme or reason. This is what makes the diagnosis of ac interference situations so difficult, and makes it necessary for us to expand both the breadth and depth of our understanding of both railroad signal systems and ac power systems. The reader is hereby encouraged to read the chapters in this handbook on railroad signal circuits (Chapter 4) and power systems (Chapter 3). In order to effectively diagnose ac interference problems, we must look at the combination of the railroad signal equipment and the ac power system as a single large system. Only by looking at this system as a whole, and examining the relationships between the various parts of the system, will we be able to fully comprehend the causes and cures of the problem. In this chapter, well focus on the causes, and will explain some of the concepts and methods that can be used in an effective diagnosis of ac interference problems. Although this represents the synthesis of the techniques of many different individuals who contributed to this work, it should in no way be construed as the only way of doing an investigation. Rather, this handbook should be used as a reference text, to be incorporated into your own diagnostic methodology as the circumstances warrant.

9-6

Investigation

You may discover some new concepts, tests, and techniques in this chapter, or may only find refinements of things that you have known for years. You may already have developed other valid diagnostic procedures of your own, which are not discussed here. The art and science of diagnosing ac interference problems in railroad signal equipment is always evolving, and there are still many needed contributions. This chapter represents a compilation of many of the currently-used diagnostic techniques whose effectiveness has been established by experts in this field over the past several years and decades. Hopefully, the documentation of these methods will provide some common ground (no pun intended) for the Railroad Signal and ac Power industries, in order to facilitate better communication and cooperation when resolving ac interference problems. Troubleshooting Fundamentals The Scientific Method Regardless of what name we give to it, the process of troubleshooting is essentially an applied version of what engineers and scientists refer to as the scientific method. One description of the scientific method goes something like this: 1. Recognize or identify a problem or question 2. Study the available information about the problem or question1 3. Guess at a possible cause or answer 4. Predict the consequences and implications of that cause or answer 5. Devise and perform an experiment to test those predictions 6. Form the simplest possible theory that explains your guess, your prediction(s), and the experimental results. This is not a new concept, with Galileo generally being credited as the Father of the Scientific Method in the 1600s. It has been revised and restated countless times since, but the basic principles remain essentially unchanged. Therefore, starting with step #1 (above), our first task is to accurately identify the problem. Recognizing AC Interference Problems When interference has occurred between ac power systems and railroad signaling systems in North America, it has historically done so in one direction only. That is, power systems have interfered with the operation of the railroad signaling, and not the other way around. With respect to the railroad operations, this is the exact opposite of the situation in the U.K. and Europe, where induction from the propulsion currents supplied to the electrically-driven trains

This step has been added to reflect modern scientific practice, and was not actually a part of the earliest descriptions of the scientific method.

9-7

Investigation

are a source of the interference, and other circuits such as telephones and other communications equipment, are the receptors. Therefore, some part of the low-power signaling equipment used by North American railroads will generally be what suffers the effects of ac interference. This is often described variously as the receptor or receiver of the interference, or sometimes the victim circuit. We should keep in mind that the most addressable cause of the problem is not always located at or near the same point in the system where the most-visible symptoms appear. However, this receptor circuit will still be where our search begins. We will then work upstream from there. In order to know when a signaling system is not working properly, we must first understand how it normally functions. Railroad signaling systems use the laws of physics and the principles of electrical engineering to automate the process of safely moving trains around a railroad. This involves the use of measurement or detection systems such as track circuits, digital logic, communication systems, and some method of displaying an indication as to when movements by trains and/or highway automotive traffic may safely be made. The reader is hereby encouraged to read the chapter in this handbook on railroad signaling circuits (Chapter 4) in order to become familiar with this equipment. Generally speaking, the effect of ac interference on railroad signaling systems is to prevent the correct transmission, reception, decoding, or interpretation of the relatively low-voltage and lowcurrent electrical signals used by the railroad signaling equipment. When a safety-critical signaling system is unable to perform its intended function, the system detects this condition and takes the appropriate action to preserve the safety of the users of the system, a right-side failure is said to have occurred. A wrong-side failure has occurred when a safety-critical system is unable to perform its intended function, and takes no action to protect the safety of the systems users. Wrong-side failures can have grave, expensive, and even fatal consequences, and are taken very seriously by the railroads, signal equipment manufacturers, and electric utilities due to the liability exposure that they represent. In more severe cases, the interference may be of sufficient magnitude to inflict permanent damage to railroad signaling equipment. This can prevent the railroad signaling system from working correctly until it has been repaired, and may result in either right-side or wrong-side failures. Railroad signaling equipment, which is designed to improve the safety of railroad operations, is intentionally designed to exhibit only right-side failures. It does this by continually checking the safety-critical components, and making sure that all of the critical electrical signals and processes, both internal and external to the system, are within their proper bounds and functioning correctly. If they are not, a right-side failure mode is entered. A right-side failure mode is a state, condition, or mode of operation wherein the system takes the appropriate action, or series of actions, to protect the safety of the users. This action often takes the form of having wayside signals turn red (stopping the trains), or having the warning devices at grade crossings activated, even when there are no trains around. In some cases, this right-side failure behavior is triggered by a physical failure of some internal or external component. The equipment detects this failure, and takes the necessary steps to 9-8

Investigation

provide continued safety. But in the case of ac interference, all of the internal and external components may still be working exactly as intended, and yet the right-side failure mode has still been triggered. (Refer to: Abnormal Operation, Chapter 5.) However, this right-side failure design philosophy creates a situation where even small amounts of ac interference (if they are of the correct magnitude, frequency, phase, modulation, or timing, etc.), can be viewed by the railroad signaling system as cause for entering a right-side failure mode. This is often the first symptom of ac interference on a railroad. The equipment used to control railroad/highway at-grade crossings (usually known simply as grade crossings) is among the most sensitive of all of the different types of signaling equipment used on a railroad. Therefore much of our discussion will be focused on the Motion Sensors, Crossing Predictors, and other audio-frequency track circuits used to control the warning devices at grade crossings. Unfortunately, there are a number of potential causes for a railroad signaling system to exhibit right-side failure behavior, most of which have no connection to ac interference. As the first visible symptoms are identical, we must investigate further in order to determine whether or not ac interference is playing a role in each case of reported problems. Obviously, the failure of the signaling system to perform as expected is the basis of most trouble reports submitted to a railroads signal department. When a deviation from normal operation is noticed, and this is reported to the responsible railroad maintenance personnel, then it becomes a problem to be resolved by the signal department, sometimes in conjunction with other departments or agencies. So this is where our troubleshooting process really begins: with a report of the problem. Step 1): Recognition A Report of a Problem In order to pursue the cause of a problem in a scientific fashion, we must follow the basics of the scientific method. However, the scientific method is only an outline, and lacks many details. In order to carry out each step of the scientific method, many other smaller steps must be taken. In many cases, these smaller steps follow a similar pattern. Therefore, we will pause briefly to introduce a memory aid that may be useful as we follow each step of the scientific method in this chapter. This sub-process is simple, having only three steps: a. Question, b. Examine, c. Decide; or Q.E.D. for short2. Our Q.E.D. is simply a way of reminding ourselves of the three most fundamental steps of the troubleshooting process: Question, Examine, and Decide.
2

The mathematically inclined may try to confuse this with the Q.E.D. statement (Latin for that which was to have been demonstrated), which is sometimes used at the end of a mathematical or logical proof. But they are different.

9-9

Investigation

The Question part of Q.E.D. begins at the time of the first report of trouble. This is when we put on our detective hats and start asking the basic who/what/where/when questions like: 1. Who reported this? 2. What happened? 3. Where did it happen? 4. When did it happen? 5. Is it still happening? 6. When did it stop? 7. In what way was this different than the expected behavior of the system? 8. What should have been happening at that time? 9. Has this ever happened before? 10. Have there been any already-identified causes of this unusual behavior? 11. Is there anything unusual taking place that may have caused this? 12. What has changed on the railway or power system? Obviously, the list of possible questions we could ask is endless. But a good investigation of a reported case of ac interference begins with these basics. This is how we begin to carry out the first step of the scientific method (see above), and identify the question or problem. This is also the first place we can apply the Q.E.D. process. We ask these basic questions (and more), examine the answers, and decide what the answers we have gotten really mean. We must also remember that examining the answers often involves looking at the amount of information available and determining whether or not we have enough information to make any decision at all. If not, more questions must be asked and answered until a decision can be made. The first major decision made during the investigation must always be: Did something really happen? If there is a credible report, which suggests that the behavior of the railroad signaling system was significantly different than it should have been at that time, then we must conclude that something unusual happened. It may or may not have anything to do with ac interference, and may or may not indicate the presence of a problem in need of correction - but once we have established that an unusual event which could affect the safety of the system in any way did occur, we are duty-bound to make a good faith effort at finding the cause. Step 2): Study the Problem Q.E.D. is also how we perform the second step of the scientific method. Once we have determined that something unusual really did occur, we must then find out as much information as possible about it. This is where the detective work begins to fan out in many different 9-10

Investigation

directions, and we will begin by casting as wide a net as possible, in order to learn what information may be available. System data recorder reports, eyewitness accounts, and any measurements taken at the time of the event (or later) must all be collected for analysis. This evidence will be used to help us make the second major decision in the investigation, i.e.: What type of deviation from normal system behavior was observed?. At this point we are still not concerned with the causes. For now, focus only on the symptoms, i.e. the precise nature of the unusual behavior of the system. We simply need to begin with a clear definition of the anomaly. This step of identifying the problem is essential to the success of all of the steps that follow. However, it is seldom given the respect and time that it deserves. Answering all of the questions that need to be asked before we proceed can require minutes, hours, days, weeks, or even months of effort. If this sounds excessive, remember that we may be attempting to find causes as small as a sliver of metal across an insulator, in a system that may extend for hundreds or thousands of miles in several directions. Only by accurately identifying the nature of the symptoms can we begin to reduce this troubleshooting problem to a manageable size. Other things to consider during this fact-finding stage are whether there are other locations that have similar symptoms with which we have dwelt in the past. Did we review trouble reports of the railway and power company for this area and what information can we glean from them? We can take some preliminary rail-to-rail and rail-to-ground voltage measurements as described in the first part of Chapter 8 and Chapter 10. Step 3): Guess at the Cause(s) When we decide that we have enough details about the problem to make an intelligent guess (called a hypothesis) at the cause(s) of the problem we do so. Sometimes we may come up with several competing hypotheses which each must be tested. This is done in order to focus our investigation into a limited number of directions. The process of making intelligent guesses at the cause(s) of the problem is perhaps the most intellectually demanding part of the entire investigation. It requires detailed knowledge of the inner workings of the systems involved (e.g. railroad and power systems), critical thinking skills, and imagination. The greater the amount of each the better. It is these same three requirements that make theoretical nuclear physics so challenging. Anyone can guess at the cause of a problem. The difficulty lies in making intelligent guesses. Often the best ways to come up with good guesses is to get experts in each related area of expertise together to discuss the problem from all sides. This means that railroad signal, track, and operating departments must work together with power generation, transmission, and distribution departments. If all parties are willing to teach the others about their respective worlds, and answer the questions that inevitably come up, everyone benefits. Sometimes these benefits appear in the form of good guesses as to the cause of the problem. Q.E.D. can be applied here too. We need to ask the question: What might be the cause of the problem?, carefully examine the guesses we come up with, and decide if theyre reasonable. If they are, they get added to the list of hypotheses to be tested. In railroad signal troubleshooting, this is often a key point at which serious mistakes are made. Usually, the mistakes result from not considering enough possible causes of the symptoms. 9-11

Investigation

Remember, the resulting right-side failure behavior is often similar for a wide variety of different causes. For example, when the gates are down at a railroad grade crossing with no train nearby, the list of possible causes is almost endless. However, a surprising number of reports like this have immediately been attributed to ac interference - even before a single measurement or observation has been made! Once we have assembled our list of intelligent guesses at the cause of the problem, we are ready to proceed. Step 4): Predict the Implications Now that we have one or more guesses or hypotheses, we need to turn them into something more useful. We need to refine one or more of them into a theory. The purpose of a theory is to explain the experimental data that we observe, and wrap it all up in a neat and useful package. But theories are not only used to explain the data that we have already collected. They can also be used to predict the system behavior that we expect to see in the future. In order to turn a hypothesis into a theory, we must identify the implications of the hypothesis, and then test them. If the results of these tests come out like we predicted, then our hypothesis can now fairly be called a theory. But this testing can be much more thorough, and sometimes much easier, if we first identify as many of the different ways of testing the hypotheses as possible. This means that we must think of as many implications of each hypothesis as possible. Suppose that we thought that induced ac energy from a nearby power line was causing a grade crossing to enter a right-side failure mode. Specifically, the crossing would become activated, i.e. flash the lights, ring the bells, put the gates down, tie up traffic, etc. when there were no trains anywhere near the grade crossing. If we have hypothesized a cause-and-effect relationship between the behavior of a particular grade crossing, and the presence of induced ac energy on the railroad from nearby power lines, then we should also be able to predict some of the implications of this hypothesis. For example, we might expect some or all of the following statements regarding the interference present to be true: 1. The crossing will be in a right-side failure mode when there is some measurable level of induced ac energy from the power lines present on the railroad rails, wiring, or equipment. (Continuous Problem) 2. The crossing will not be in a right-side failure when there is either no measurable level of induced ac energy from the power lines, or the level of the induced ac energy is significantly lower than the level needed to put the crossing into a right-side failure mode. (Normal Operation) 3. There is a definite correlation in time between the presence of the necessary level of interference, and the crossings right-side failure mode. Specifically, the crossing will enter 9-12

Investigation

the right-side failure mode soon after the necessary level of interference has been achieved, and the crossing recovers from the right-side failure mode soon after the necessary level of interference is no longer present3. 4. If the level of interference is not constant, there may be a range of interference levels that cause only intermittent right-side failure modes, and the frequency of such right-side failures increases with the level of the interference. 5. The level of the ac interference on the railroad shows a correlation with some characteristic or activity of the nearby power line such as changes in the phase current, phase current unbalance, switching operations, capacitor bank activity, etc. This is the step in the scientific method that requires us to flesh out our hypotheses in preparation for testing them. Beginning with our preliminary observations, and taking each of our hypotheses in turn, the logical implications of each hypothesis must be identified. These identified implications will form the basis of our tests. Step 5): Test Each Implication The testing of the implications of each hypothesis can require a large amount of time, effort, and expense. Often we will only test what we think are the most definitive implications or predictions stemming from a given hypothesis. That is, we will only perform the tests that will most quickly discriminate between the various competing hypotheses, and identify those that correctly explain the data. However, we must also consider the testability of each prediction. For example, if we think that induced ac energy from nearby power lines is causing a problem, then there are a number of ways in which this might be tested. The most direct way of testing this would be to shut down the power lines. Obviously, this is not always practicable, and not often used. Therefore, we must look for other ways of testing this hypothesis. One alternative might be to apply test shunts across the tracks near the affected signal equipment, so as to shunt the interfering ac voltages seen by the signal equipment. By installing a direct and low-impedance connection from one rail to the other, both rails should now be at the same potential. If a voltage was previously being induced onto the rails in an unbalanced fashion, then the test shunt(s) would tend to equalize this voltage. Since most railroad signal equipment only senses the difference in voltage between the two rails, and is only subject to ac interference that causes an unwanted difference in voltage between the two rails, we would expect to see the behavior of the signal system change when we applied the shunt(s). We would probably expect the signal equipment to leave its right-side failure mode, and return to normal operation (depending on where and how the shunt(s) was/were placed).

Most crossing warning equipment such as Crossing Predictors and Motion Sensors have built-in delay timers that may keep the crossing activated for a specific period of time after the interference has dropped below the level needed to cause a right-side failure mode. Often these timers are roughly 30 seconds in duration.

9-13

Investigation

However, placing a shunt across the tracks affects many things, and not just the ac voltage balance of the rails. Therefore we might look for some other less-invasive method of testing. We might choose to measure the ac voltage from each rail to ground with a voltmeter, as well as the rail-to-rail potential, over a period of time. If ac induction from the power lines really were the source of the problem, then we would expect that the voltage induced into the rails would vary as a function of the current or current unbalance on the power lines4. Furthermore, we would expect that the failure rate of the signal equipment would be a function of the interference level on the tracks, specifically the rail-to-rail ac voltage(s) at the power line frequency, or at one of its harmonics. The signal equipment might exhibit a sort of threshold behavior, where the failure rate (measured in false crossing activations per hour, or something similar) would change sharply above a certain threshold level of measured rail-to-rail voltage, or measured current/current-unbalance on the power lines. Assuming that the load or degree of unbalance on the power line changed over time, and that the needed measurements could be taken by both the power company and the railroad signal department, and that the power demand changes enough (daily, seasonally, etc.) to make the railroad equipment both misbehave, and work normally, this could be an effective way to test for a connection between the characteristics of the power line, the level of the ac interference on the railroad, and the behavior of the railroad signal equipment. In this example, we expanded our original hypothesis: induced ac interference from the power lines is causing the problem, by looking at the implications of this hypothesis. Specifically, we looked at the level(s) of ac power line frequencies present on the tracks, to see if it correlated with the characteristics of the power being carried by the power line. We also looked to see if the level of ac interference on the tracks had any correlation with the false activation rate at the grade crossing. Both of these tests are based on the implications of our original hypothesis. There are many other ways in which this relationship between the power lines and the signal equipment might be tested. But the point here is that our hypotheses must lead to predictions, which can be turned into test procedures that can be carried out. Step 6): Form the Simplest Theory that Explains the Results. Once we have conducted a test, we must then decide what the results mean. If our results confirm our hypothesis, then we would tend to accept it as a working theory, and either go on to look for more ways of confirming or disproving it, or proceed to apply the theory in order to find a solution to our problem. Remember: Scientists study things, while Engineers study things and then make them work. Being sure that you have enough data to confirm your hypothesis, and then proceeding to apply your newly developed theory is often difficult. This decision is usually a judgment call, hopefully based on past experience.

This assumes that the ballast resistance from rail-to-earth is essentially constant. The voltages measured may be affected significantly if the ballast resistance changes substantially over the test period.

9-14

Investigation

If our results do not confirm a hypothesis, then we must abandon it and test another hypothesis. Often cases of ac interference are correctly diagnosed only after many other hypotheses have been discredited. Unfortunately, the outward symptoms of ac interference are very similar to many other problems. Therefore we must often test for a large number of other more mundane causes before concluding that ac interference is the source of the railroad signal equipment problem. A Hidden Fork in the Road We have now arrived at a hidden, and all-too-often overlooked, fork in the road as we head towards a solution to reported ac interference problems. This handbook was intended primarily as a guide for the diagnosis, detailed identification, and successful mitigation of ac interference problems in railroad signal equipment. From this point onwards, we will make the general assumption that the source of the problems being experienced by the railroad signaling systems is interference from ac power systems. However, this is not always the case, and there is another potential solution to our railroad signal equipment problems that must be briefly but seriously considered. A wise man once said: When you hear hoof-beats, expect horses - not zebras. This is one of the fundamental tenets of troubleshooting. In other words, we should look for the common and ordinary causes of trouble first, before we resort to less-probable hypotheses. Many signal system problems are too-quickly attributed to ac interference, high-tension power lines, or induction when in fact the trouble has actually resulted from ordinary signal equipment failures or intermittencies. Even in cases where the ac power system is found to be playing a role in causing the symptoms observed, the root cause can still be a garden-variety failure of a single component such as a surge protection device, filter, or insulator. The physics of ac induction as it applies to railroad signaling is unfortunately not as well understood by railroad maintenance professionals, as a whole, as we may wish it were. Often induction has been the explanation of last resort for a weary and haggard signal maintainer at the end of his troubleshooting rope. But this desperation does not relieve us of the responsibility for verifying that all is as it should be with regard to the health and general condition of each signaling system component before launching into an investigation of interference from the ac power system. Before we begin to analyze the problems exhibited by the signaling system with an eye towards possible ac interference, we first need to have determined two things: 1. We need to have verified that there is no underlying failure in the signal equipment that would have been detected by standard railroad signal maintenance and troubleshooting practices, and that all components of the signaling equipment are in good working order. 2. We need to have positively identified the specific equipment or system that is causing 5 the observed problem .
5

We need to have positively identified the specific abnormally functioning signal equipment component or subsystem that is furthest upstream in the chain of control connected to the apparently malfunctioning signal system appliance. There should be no problems found in the next component or subsystem upstream, and all components or subsystems downstream should show some sign of abnormal operation while the problem is occurring.

9-15

Investigation

For example, if the warning devices at a grade crossing are activated when they shouldnt be, then it is probably not the lights themselves that are functioning abnormally, but rather the devices supplying power to, or otherwise controlling, the lights. In many cases, the devices supplying power to the lights are not at fault, but rather the equipment controlling these devices. It is unproductive to blame as-yet-unverified ac interference for unwanted operation of the lights and other warning devices at a grade crossing, when the root cause is actually a blown powersupply fuse in the Motion Sensor or Crossing Predictor controlling the crossing. If we look at all of the different causes of railroad signal equipment trouble, ac interference accounts for only a very small fraction. Mundane causes such as failed components are much more likely. Only when we are sure that there are no un-diagnosed problems with the railroad equipment itself should we proceed to look for ac interference. As we discuss the process of investigating and diagnosing cases of ac interference affecting railroad signal equipment, we will frequently point out other more ordinary, and often more likely, causes of the same symptoms. The Source Path Receptor Model The scientific method, as described above, is a very powerful tool for learning about our physical world. It does, however, have a few drawbacks. For instance, the number of possible hypotheses and potential implications of these hypotheses is nearly infinite. Actually testing all of them would take forever. We dont care about the great number of incorrect hypotheses that we can form. We are only interested in the hypotheses that can be confirmed by testing and experimentation, and can be developed into theories explaining the nature of the problem. Fortunately, in the case of ac interference, all of the hypotheses and theories follow a single pattern. This pattern is called the SourcePathReceptor model. The Source-Path-Receptor model is a model used by Electromagnetic Compatibility (EMC) engineers to identify the three essential components of any EMC problem. In every case of problems with EMC, there are always three things present: 1. A Source of interfering electromagnetic energy, 2. A Path for the interfering energy to follow from the Source to the Receptor, 3. A Receptor that is affected by the interfering electromagnetic energy. We use the scientific method to guide our troubleshooting process. Troubleshooting ac interference problems in railroad signal equipment consists of identifying the component or components that are fulfilling each of the three roles listed above. We do this by progressively reducing the size of the physical area we are searching until we have isolated a single component or single set of components. Each time we test a hypothesis, it is with the intent of verifying or eliminating a hypothesis associated with a particular component. 9-16

Investigation

Mitigation or elimination of an ac interference problem consists of modifying the characteristics of the components in one or more of the following three categories. Our objective in mitigation is to: 1. Reduce the strength of the Source, 2. Reduce the transmissive efficiency of the Path, 3. Reduce the susceptibility of the Receptor. The source of the interfering electromagnetic energy in most instances will be some portion of the ac power system. This may be a portion of the nearest ac power line, a transformer, a capacitor, a substation, or even the entire power system leading back to a power generation plant. The path for the electromagnetic energy will include everything between the source and the receptor, which acts to carry the ac energy by conduction, induction, or radiation. The receptor will generally be the affected railroad signal equipment. This model is very simple, and very important. The Rules of Thumb of Railroad Signals and AC Interference Now that weve tipped our hat to Galileo, the Scientific Method, and the Source-Path-Receptor model, its time to get down to basics. These are the basic facts and prevailing opinions that should guide your investigation of a suspected case of ac interference with railroad signal systems. Keep them in mind when conducting any investigation of suspected ac interference. 1. Ninety percent of everything is track. This means that of all of the potential sources of trouble in railroad signaling, something in, on, or very near to the track itself is the source of trouble in about nine out of ten cases. The sophisticated electronic equipment located inside the equipment shelter (case or bungalow) is really pretty reliable, provided that it has been installed and maintained correctly. In the case of ac interference, the track is normally at least some part of the most common paths for the interference. 2. Motion Sensors and Crossing Predictors are the most sensitive devices connected to the track, and will usually be the first indicators or victims of ac interference. Of the two, Motion Sensors are the more susceptible. Other types of railroad signal equipment may also be affected, but the level of interference usually must be higher in order to cause problems for these other systems. Normally the first symptom of ac interference with crossing equipment will be false activations of the warning devices at a crossing. 3. Railroad signal circuits see the rails of the track differentially. This means that most railroad signal circuits can only sense the difference in voltage from one rail to the other on the same track. In theory, if both rails were raised to an electrical potential of hundreds of volts (ac), this would have no effect on the operation of the equipment provided that both rails were of the same elevated voltage and phase. With few exceptions, only voltages that appear between the rails can affect the operation of the signal circuits. 9-17

Investigation

4. Anything which unbalances the electrical characteristics of one rail with respect to the other can act as a catalyst in turning induced common-mode voltage into rail-to-rail voltage. If the two rails of a railroad track are both in the same electromagnetic field, then they will normally have the same voltages and currents impressed into or onto them. But railroad rails lie on a less-than-perfectly-insulating medium known as ballast (the crushed rock used by railroads). If one rail is effectively connected to ground (remote earth) to a greater degree than the other, then there will be a difference in the voltages of the two rails. Shorted lightning arresters in the railroad signal equipments so-called surge protection equipment are famous for doing this. Shorted insulated joints in the tracks and failed insulators on concrete or steel ties can also do this. 5. Generally speaking, a lot of cases of induction in reality turn out to be conduction. Although ac power lines can create an electromagnetic field, and railroad tracks or wires can turn that field into induced ac currents and voltages, this isnt the most common way of winding up with an ac voltage difference between the railroad rails. Unintentional paths that connect one rail or the other to a source of voltage can create serious and even equipmentdamaging rail-to-rail voltages on the track. Shorted railroad lightning arresters are a good way to make this happen. Since the ground bus at many signal equipment installations is connected to the Neutral bus of the ac power service to the location, these installations are often only one shorted lightning arrester away from connecting one rail directly to Neutral. On highly-unbalanced power lines, or in areas where the distribution pole grounds are damaged, missing, or ineffective, the few volts of ac energy present on the Neutral wire may be enough to cause problems for the signal equipment if the Neutral is connected directly to one rail via a shorted arrester. Other local wiring problems can also create a hot Neutral bus in the signal equipment enclosure, but an unintentional conductive path to one of the two rails is still essential in order for problems to occur. 6. Just because theres some ac interference on the track, this doesnt mean that ac interference is the cause of the problem. In order to interfere with the normal operation of signaling equipment, the induced ac energy may need to be of the correct amplitude, frequency, phase, modulation, etc. A low level of induced ac energy is normally present on most railroad tracks. However, its amplitude is low enough that it has no effect on the operation of the signal circuits. For Motion Sensors and Crossing Predictors, a quick rule of thumb is to compare the level of the power line frequency or harmonic that is closest to the operating frequency of the affected signal circuit. If the two frequencies are close or adjacent to one another (roughly within half an octave of one another), then the level of the interfering ac voltage (measured from one rail to the other) generally must be greater than the track circuit signal, or less than 10 dB below it, to have an adverse effect. (Guidelines from the signal equipment manufacturer should always be consulted in any specific situation.) 7. Many railroad signal circuits are frequency-selective, but enough of even a non-adjacent frequency can cause operational problems. Obviously, if enough voltage or current is applied to the signaling equipment to cause permanent physical damage, this can affect the normal operation of the equipment. (See chapter on Equipment Damage). But below this amplitude lies a range of energies that, even though they will not cause permanent damage, will cause abnormal operation of the railroad signal equipment. This occurs because there are limits to the frequency-selectivity of the signal equipment, and because it is also possible to saturate the transmitters and receivers of the signal equipment with ac interference energy. 9-18

Investigation

8. When you hear hoof-beats, expect horses, not zebras. As stated above, the basic rules of common sense and statistics apply to troubleshooting potential ac interference problems on railroads. That is, it makes sense to check the most likely causes of a given symptom before hypothesizing about ac interference. Too many signal maintainers have given in to the temptation to declare a problem to be caused by ac interference, and simply walk away. Due diligence requires that we check all of the basics first. Is the signal equipment in good working order? Have all of the factory-recommended updates been installed? Has the equipment been well cared for? Even though ac interference can cause intermittent rightside failures at a grade crossing, a dying battery charger is more likely to cause the same symptoms. 9. The reliability of a railroad signal component may be proportional to its distance from the track. The railroad wayside environment can be a severe one. Extremes of temperature and humidity are accompanied by vibration, dust, moisture and other environmental factors known to reduce the reliability of most equipment. Rails take an incredible pounding, cyclic loading. Track circuit components that are buried directly in the ballast of the track (track wires, termination shunts, couplers, etc.) may have been damaged by track maintenance equipment (ballast tampers and ballast regulators, most notably), or may simply have failed due to the fatigue from having trains driven over the top of them day in and day out. Generally, the farther that equipment is located from the tracks, the more reliable it generally is (see #1, above). 10. Transmission lines may look big and menacing, but ordinary distribution lines are more often the source of ac interference. This happens for several reasons: a. The primary mode of induction between ac power lines and railroads is magnetic, and not electric. Therefore the magnitude of the current plays a larger role than the magnitude of the voltage. The magnitude of the currents carried by distribution and transmission lines are often more similar than one might expect. The more current carried, the more potential for ac interference. b. Distribution wires are usually physically closer to the rails or line-wires of the railroad receptor circuit than the wires of transmission lines. Since electromagnetic radiation is an inverse-square-law field, cutting the distance in half causes the strength of the electromagnetic field to quadruple. c. Placing a multi-phase ac power line closer to the receptor circuit (as with distribution lines) often makes the physical geometry of the ac power line play a larger role, i.e. the degree of cancellation between the fields created by the phase currents gets lower. Remember, the field at the rails will be the vector-phasor sum of the fields created by each current-carrying wire of the ac power line. d. Distribution lines normally have a larger amount of harmonic content than transmission lines. In railroad signaling, it is often a harmonic of the 50 or 60 Hz power-line frequency that interferes with the signal equipment. This happens for many reasons, but primarily because harmonics of the power-line frequency are normally much closer in frequency to the operating frequencies of the signal equipment. Furthermore, these higher-frequency harmonics will be more efficiently coupled into the railroad system, due to the increased mutual reactance at these frequencies (see Chapter 2). 9-19

Investigation

e. Phase current unbalance is usually greater on distribution lines than transmission lines. Greater phase current unbalance means greater residual (zero sequence) current. Magnetic fields from residual currents drop off roughly in proportion to the distance as opposed to the square of the distance (inverse-square-law) for balanced currents. This is because part of the residual current flows in the earth. These rules are not definitive, and should only serve as guidelines for your investigation. But they summarize most of the recurring patterns in ac interference cases, and can often save you time. The Questions Now that weve outlined the basic guidelines in railroad signal/ac interference troubleshooting, lets return to steps 1 and 2 of the scientific method, and get into the nitty-gritty details of investigation and diagnosis. Since the questions that we ask are so critical to our success in locating and identifying the components serving as the source, path, and receptor in an EMC problem, we have attempted to assemble a list of troubleshooting questions below that may apply to many troubleshooting situations. You will notice that many of these are also the same questions as in the troubleshooting flowchart, but here we have tried to give the background and implications of each of these questions. A list of the basic questions can be broken down into three big categories: What, Where, and When. These are the questions that form the basis of steps 1 and 2 of the scientific method outlined above. These are also the questions we can use to begin narrowing our search for the root cause or causes of the problem. The What Questions These are the first questions that an investigator always asks. They may lead in any number of directions, but the most important thing to remember when asking them is not to pre-judge any of the answers, because you never can tell where some of them might lead. In the case of railroad signaling, we usually ask specific questions relating primarily to two of our five senses: what was seen, and what was heard. But occasionally we will also ask about what was felt, smelled, or tasted. Our goal here is to gather all of the information that we can about the symptoms of the problem, which presumably began occurring at some point in the past. The what questions should include: 1. What deviation from normal functioning was observed? This is usually our very first question, usually followed closely by the question: What should have been happening instead? It is this comparison between the actual and the

9-20

Investigation

expected that sets us on our course of investigation, so we must ask it very carefully, and listen carefully to the response. 2. What equipment is directly responsible for this function? This question is our first step upstream as our investigation proceeds from the symptom(s) to the cause(s). 3. What other equipment is indirectly responsible for this function? This question represents a second step upstream, and is important because it reminds us that the symptoms that were observed may not have been caused directly by the equipment which was seen to malfunction. Instead, the very nature of railroad signaling, with its very long control chains, makes indirect causes not only possible, but likely. For example, a light bulb in a signal light that should be lit, but isnt, could be any number of things. It could be as simple as a burned-out bulb filament, but it could also be a worn out bulb socket, a broken wire, a damaged contact on a relay, a dead battery, or even a failed battery-charging system. Always remember that the cause isnt always found at the same location as the symptoms. This leads us quite naturally into the area of Where, which will be discussed next. The Where Questions Any description in response to the What questions, above, will usually contain some Where information as well. These are the questions that we ask in order to draw some geographical and spatial boundaries around the problem. 1) Where did this happen? Where were the symptoms observed? Was the problem observed to occur at more than one location at the same time? Could more than one location be observed while the symptoms of the problem were occurring? Railroad signal equipment is often distributed along the length of the railroad in similar and repeated patterns. We want to find out if other similar installations were similarly affected, and if not - why they werent. 2) Where didnt this happen? Is there a clear boundary between places where the symptoms were occurring and where they werent? As above, could more than one railroad signal equipment location be observed while the problem was occurring? Could physical boundaries be seen on all sides of the problem? Were there other tracks or signal locations nearby that were apparently unaffected? Were all of the signal equipment installations within the physical boundaries observe affected? Were they all affected equally? 3) Where/what do the circuits in question connect to? This is almost a repeat of What question #3, above. The two are closely interrelated and difficult to separate. We are trying to determine the physical scope of where the problem may lie. It is better to ask this question twice in slightly different ways, as the person investigating the problem is only rarely the person who saw it first. Relying on second-hand information, we must use our best investigative techniques to draw information out of all eyewitnesses, if there are any. Again, we are trying to draw some physical boundaries around the problem.

9-21

Investigation

The When Questions Although the general answer to each When question is usually: sometime in the past, the details are extremely important. These details are often what helps us establish cause-and-effect relationships, and characterize the fundamental nature of the problem, i.e. intermittent vs. continuous, random vs. predictable, and periodic vs. non-periodic. This helps us to draw temporal boundaries around the problem. 1. When did it start? In the absence of automated recording systems (data recorders, black boxes), it is often difficult to establish exactly when a problem occurred. However, a general idea of the timing will often suffice, and both continuous and event-driven data recorders are becoming increasingly widespread on both the railroads and electric utility systems. 2. What else was happening when it started? This is the basic question that we ask in order to begin to look for causal relationships between the symptoms and any potential causes. In railroad signaling, most failures have very little delay between the occurrence of the failure or problem, and the onset of symptoms. The danger here is that all too often, well-meaning investigators draw the wrong conclusion. For example, it is widely recognized that the rate of mortality by drowning is highest during the months in which ice cream sales are the highest, but this does not mean that ice cream causes drowning! Similarly, a crossing warning system that malfunctions only when a train approaches the crossing does not mean that the train is the cause of the problem. But be sure to get a good idea of any other seemingly unrelated events occurring on the periphery of the situation that may have been happening while the symptoms of the problem were apparent. Some of them may be relevant. 3. When did it stop? Not all problems are continuous. Sure, the ones that are continuous will generally be much easier to locate and solve. Unfortunately, interference from ac power systems isnt always this way. Getting an idea of the duration of the failure may provide clues as to its cause or source. 4. What else was happening when it stopped? As in 2), above, we are looking for causal relationships between the symptoms and other externally observable events. 5. Could it have been happening before you noticed it? We are always tempted to think of our own observations as correct, but we also need to recognize our limitations as observers. Recognizing the limitations of the information that we use in our investigations can help us weigh and interpret each piece of data correctly. 6. Was the problem continuous or intermittent? In cases of interference, this piece of information can help us identify both the potential sources, as well as the rough magnitude of the interference affecting the operation of the railroad signal equipment. For example, a level of interference that is just barely adequate to induce a right-side failure mode may not be able to do so continuously. This is often the case with crossing warning systems that are seen to activate the warning devices (e.g. lights, gates, bells) at random intervals for varying lengths of time. 9-22

Investigation

7. Did you have the problem under continuous surveillance? Again, we are trying to establish the quality of the information that we are working with. Many symptoms thought to be frequent or continuous have been found to be infrequent or intermittent once they were closely monitored. In some cases, this difference can be critical in forming a diagnosis. 8. Could it have stopped and started again while you werent watching? Sometimes this is how an investigator discovers the presence of a data recording system or other form of monitoring which was not mentioned earlier. Just because a symptom was no longer under direct visual observation does not mean that it was not being monitored. For example, if we want to know if an activated warning system at a grade crossing has recovered completely to its un-activated state, we may only need to listen for the bells. 9. Has there been a history of these sorts of symptoms at this location? This is the question that often opens up the biggest can of worms. Many times, an investigator will be flooded with the history of every problem of this general nature that has ever occurred on a particular railroad, as well as the complete history of all problems, similar or not, that have occurred at this particular location. For every kernel of wheat, much chaff must be separated. Even though it is sometimes arduous to slog through, this information should not be discarded out-of-hand, as it often provides valuable insights into the frequency, duration, and possible causes of the problem. Guessing at the Causes The third step in our scientific method is to hypothesize about the possible causes of the problem. These hypotheses should arise from the mass of information that we have gathered thus far. Provided that we have eliminated most of the ordinary causes of signaling system trouble, we may now direct our thinking toward identifying the specifics of our ac interference problem. Fortunately there is a model that describes each of the pieces involved in such cases of interference, and the characteristics of each. Electromagnetic Compatibility (EMC) engineers spend much of their time doing just this, identifying the pieces that fill each of the three roles in the Source-Path-Receptor model. The Source-Path-Receptor Model states that in order for interference to occur, we must have at least these three things: 1. A Source of Interfering Energy: In most of our cases, this will be some portion of the ac power grid in North America. However, we will usually try to narrow this down considerably. 2. A Path to carry the Interfering Energy from the Source to the Receptor: Generally, this will take one of two forms: induction or conduction. 3. A Receptor which is susceptible to the induced or conducted interference. Remember, the Source-Path-Receptor model must always be satisfied in order for ac interference to cause signaling system problems. Therefore every hypothesis that we construct in an attempt to explain the basis of an ac interference problem must provide at least one component or mechanism for each part of the Source-Path-Receptor model. 9-23

Investigation

The Specific Questions After asking our general questions, there still remains a whole host of specific questions that should be asked, or at least considered, before any diagnosis is made. These questions are all refinements of the basic what, where, and when questions - mostly the what kind. The answers to these will be our guides as we consider each possible candidate for the roles of Source, Path, and Receptor. Even though the Source-Path-Receptor model appears to begin with the Source, it is usually applied to troubleshooting in reverse order, starting at the Receptor. As indicated above, this receptor will probably have already been clearly identified before we begin our investigation of ac interference. Common Sources for E.M. Fields Power Lines If induction is the path of the interference, then things that produce electric and magnetic fields will be the sources. Anything that has a voltage on it produces and electric field and is a potential source of interference. Anything that carries electrical current produces a magnetic field and is a potential source of interference. From this description it is clear that electric power lines and substations can be sources of ac interference. Additionally, it is important to remember that electrical systems beyond the control of the power company can also be sources. This includes commercial and industrial facilities as well as residential ones. Occasionally, railroad electrical equipment has been the source. One aspect of sources to consider is whether the source is localized or distributed. In cases of magnetic induction the source is usually distributed over the length of the conductor parallel. The energy magnetically coupled into the receptor circuit is roughly proportional to the length of the parallel (i.e. longer parallel results in higher induced voltage). An example of a localized source of magnetic induction is when a locomotive with 60 Hz cab signaling passes under a transmission line crossing the tracks, and has a momentary drop-out of the cab signal system. In this case the pickup coil circuit can be directly affected at the power-line crossing. The effects of electric induction can also be either localized or distributed. Large conductive objects like trucks and rolling stock can develop static charges from either parallel or crossing transmission lines, because this is a localized phenomenon. High impedance communication circuits can develop significant voltages from parallel transmission facilities a distributed source. It should be noted that placing power lines underground effectively eliminates electric induction, but not magnetic induction. Some underground installations incorporate design practices that reduce magnetic fields, but the earth itself does little to reduce magnetic induction. Underground power lines need to be modeled including unbalanced phase currents and zero-sequence residual current in the shields and pipes.

9-24

Investigation

Common Sources for Conducted Interference Any electrically energized equipment can be a source of conducted interference under fault conditions. The fault creates an unintended conductive path. Sources of conducted fault currents can be any voltage and from any system (i.e. power company, railroad, third party). The duration of conducted fault current interference can be very brief as with high speed relaying on transmission lines. Or, the fault can last indefinitely, as with defective insulation on a power cord with enough path resistance to prevent the fuse from blowing. Under normal conditions, the number of sources of conducted ac interference is smaller. Only intentional conducted paths are present. Distribution lines with multigrounded neutrals are the most common possible source of conducted interference. Neutral current returning to the substation transformer will flow partly through the earth. And whatever current flows in the earth will get into the rails through the ballast. The density of these neutral currents tends to peak near the source substation. So, conducted levels of neutral ac interference will tend to maximize where distribution substations are located next to railroad tracks. The other possible source of conducted ac interference is energy conducted into the equipment enclosure via the phase conductors of the ac power service to the signal location. This path of interference through the electric meter is usually in the form of harmonic interference. This through-the-meter interference falls within the field of power quality, and is usually fully addressed by power company personnel with established procedures and equipment. Common Paths Fields, Rails, and Wires In our understanding of the source-path-receptor model, the path consists of everything between the source and the receptor. The exact boundaries of this are sometimes difficult to define. For our purposes, we will define the source-path boundary as lying immediately in front of the first un-mitigatable source of ac interference encountered as we move from the receptor to the source. The boundary between the path and the receptor will be defined as being located immediately in front of the first component or process whose normal functioning is affected by the ac interference as we move from the source to the receptor. In cases of ac interference with railroad signal equipment, the path often consists of the wires of the ac power system, an inductive or conductive path from the wires to the rails, the rails themselves, and the track wires connecting the rails to the signal equipment. In some cases, 6 the rails will not be a part of the path, as in cases involving open line-wire or above-ground cable on the railroad. The characteristics of the path for any inductively-coupled ac interference depend primarily on the physical geometry of the ac power line and the railroad rails or wires. The distance between
6

Open line-wire is the wire strung on the telegraph poles often found running next to older railroad lines. It is rapidly disappearing from North American railroads, and is being replaced by other means of communicating the occupancy of various sections of track up and down the railroad line.

9-25

Investigation

the power line and the railroad, the arrangement of the power lines conductors, and the length of the exposure between the power line(s) and the railroad track(s), all play a role. Normally, any electromagnetic field created by a power line will induce some measurable voltage and current onto nearby conductive objects. Given the shape of railroad rails, this induction is usually magnetic in nature, and affects both rails of the railroad track equally. As most of the signal and track circuits used on a railroad are (quite intentionally) not referenced to remote earth, and use differential voltages to detect trains and communicate information, this equal induction onto both rails of a railroad track has no effect on the operation of the signal equipment. However, if there is a significant difference in the electrical characteristics of the two rails in a section of railroad track, there can also be a significant difference in the rail-to-ground potentials induced onto them by the electromagnetic field from the ac power line. This difference results in some portion of the induced rail-to-ground voltage potential appearing as rail-to-rail voltage potential. Any imbalancing agent on the railroad, such as a shorted insulated rail joint, a shorted lightning arrester, unequal ballast resistance, or any other unequal connection to ground or other conductors in the same electromagnetic field can act as a catalyst for converting induced 7 8 common-mode voltages into differential-mode voltages . In this way, the rails and wires of a railroad can sometimes act as more than just a simple path for ac interference. In some cases of ac interference, the inductive coupling via an electromagnetic field from the power lines to the rails or wires of the railroad plays a relatively small role. In these cases, the primary path for the ac interference is an entirely conductive one. Return currents in the earth may connect to one rail of the track more than the other, through unequal ballast resistance from each rail to ground. And in places where the voltage gradient in the surface of the earth is high, as it is near concentrations of earth currents such as power generation facilities and substations, an interfering differential ac voltage may appear between the rails even if they both have the same rail-to-ground resistance. In some cases the path for the ac interference bypasses the rails almost completely, and enters the signal equipment directly through an unintentional connection to a source of ac power. This can occur when either the Hot or the Neutral wire of the commercial power service to a signal equipment installation becomes connected to the local ground bus of the railroad lightning protection equipment (lightning arresters and equalizers). If there is an undiscovered lightning arrester present which has failed in a shorted manner, this can provide a direct connection from the ground bus to one rail of the track.

7 8

Energizing both rails equally with respect to remote earth potential. Energizing each rail to a different ac potential with respect to remote earth.

9-26

Investigation

Common Receptors Motion Sensors and Crossing Predictors Motion Sensors and Crossing Predictors Symptoms and What They Mean The most common piece of railroad signal equipment to be affected by ac interference is the Motion Sensor, followed closely by the Crossing Predictor. Motion Sensors and Crossing Predictors are alike in that ac interference will generally put the crossings warning devices into operation when they shouldnt be. In order to understand how this can happen, we will begin with a review of the basics of Motion Sensors and Crossing Predictors. Motion Sensors and Crossing Predictors work by applying a single-frequency (sinusoidal) constant-current ac frequency in the range of 40 Hz to 1 kHz to the track (the operating frequency), and measure the voltage that results. The magnitude of the resulting voltage will be proportional to the impedance of the track. The impedance of the track is limited by installing termination shunts across the track some distance away from the grade crossing itself. This distance is determined by the maximum speed of the trains, and the amount of warning time desired at the crossing. The sections of track between the grade crossing, and the termination shunts on either side of it are referred to as the approaches of the crossing. These will range in length from a few hundred, to a few thousand feet in length. When there are no trains within the approaches of the crossing, the impedance of the approaches will be at a maximum, and so will the magnitude of the ac voltage at the operating frequency of the Motion Sensor or Crossing Predictor. As the train approaches, the impedance of the track will diminish, as it is shortened in length by the shunting action of the first axle of the approaching train. The magnitude of the ac voltage at the operating frequency of the Motion Sensor or Crossing Predictor is generally proportional to the distance of the train from the crossing, and the rate of change of the magnitude is proportional to the trains velocity. Motion Sensors activate the crossings warning devices whenever the speed of the approaching train (as measured by the rate of change) is above some threshold. Crossing Predictors activate the crossings warning devices at an essentially constant interval ahead of the trains arrival at the crossing. They do this by comparing the magnitude of the operating frequency voltage, with its rate of change. This follows from the familiar Time = Distance/Velocity equation. Whenever the time computed is less than the desired constant warning time, the warning devices at the crossing are put into operation. Both of these processes of train detection can be adversely affected by ac interference. The crossing will remain activated until the train passes through the crossing, or has stopped short of the crossing for several seconds (usually 15 to 99 seconds). In order to determine when the tail of the train has cleared the crossing, a separate dc or audio-frequency track circuit known as an island circuit, usually built into the Motion Sensor or Crossing Predictor, checks the track for occupancy in the immediate vicinity of the grade crossing.

9-27

Investigation

Motion Sensors, Crossing Predictors, and almost all track circuits using audio frequencies are connected to the rails of the railroad track using some form of coupling network involving a capacitor, and a transformer. Normally, there will be separate coupling networks for the transmitter that applies the operating frequency to the track, and the receiver (or receivers) doing the detection. The transmitters coupling network will usually be of lower impedance (a few Ohms, or tens of Ohms) than the receivers coupling network (tens of Ohms, to thousands of Ohms). In order for ac interference to affect the operation of a Motion Sensor or Crossing Predictor, it must be present in the form of a differential voltage. That is, the interfering ac voltage must exist as a potential between the two running rails of the same railroad track. Otherwise, it would never be able to pass through the coupling network of either the transmitter or receiver of the Motion Sensor or Crossing Predictor. Once significant ac interference has become present on the rails in a differential fashion, the symptoms will usually take one of several forms (see below). We will examine each common set of symptoms and review the possible causes. For the sake of brevity, the terms Motion Sensor and Crossing Predictor will be abbreviated as MS/CP.
Case #1: The Crossing Equipment is Activated Continuously or Intermittently When there are No Trains Around

Intermittent activation of grade crossings with no trains in the area is the most common symptom of ac interference. There are five basic methods for ac interference to cause these types of operational symptoms. 1. Spoofing of Train Motion. If the ac interference is close enough in frequency to the operating frequency of the MS/CP for a significant amount of interference to make it past the receivers filters (significant = at least a few percent of the operating frequency amplitude), then the accuracy of the measurement of the magnitude of the operating frequency voltage by the MS/CP can be affected by any fluctuation in the magnitude of the ac interference. If the level of the interfering ac frequency dropped significantly over a period of a few seconds, this could appear to the MS/CP as an incoming train, even when there were no trains in the area. This would put the crossings warning devices into operation, at least temporarily. When the MS/CP no longer saw what looked like an approaching train, a timer would begin running, and the crossing would recover if no further incoming trains were seen by the MS/CP before the timer (usually 15-99 seconds in length) expired. Because they only require the appearance of motion above some threshold value, Motion Sensors are much more susceptible to this form of ac interference than are Crossing Predictors. This is sometimes called in-band interference, when the interfering frequency is very close to the operating frequency of the MS/CP. However, there is also an easier way to create this. If two MS/CPs are operating on the same track, and have been set up to use the same operating frequency, this will normally work fine, provided that the approaches between the 9-28

Investigation

two MS/CPs dont overlap, and that the termination shunts in between the two crossings are far enough apart. However, if either of the termination shunts between the two crossings fails open, then the nearly-identical operating frequency from two MS/CP transmitters will be received by both MS/CPs. This is called on-frequency interference. With modern crystalcontrolled oscillators, the two operating frequencies will normally be within 0.02% of one another, but will never be exactly the same. These two nearly-identical operating frequencies will mix together, forming a beat pattern whose frequency is equal to the difference in the two frequencies. The resulting time-varying amplitude can look like the motion of a train to either or both of the two MS/CPs. One or both of them will likely be activated intermittently, usually in a regular and periodic fashion (like every few minutes). The keys to recognizing this non-ac-related problem are that one or both of the MS/CPs will 9 10 show a fluctuating train position reading , and possibly a fluctuating track quality indication . Furthermore, these will vary in a very regular and repeated pattern. Any two MS/CP units within the same electrically-continuous length of track, and within less than a few miles of each other can potentially have this problem, if they are using the same operating frequency. 2. Simulation of an Error Condition. Although the primary purpose of Motion Sensors and Crossing Predictors is to detect approaching trains, in order to do this in a vital and fail-safe fashion (see Chapter 5, Abnormal Operation of Railroad Equipment), they must constantly test for the presence of conditions that could result in a short warning time, or no warning time. This is done by measuring several characteristics of the electrical signals applied to the track by the transmitter. The phase, frequency, and other characteristics of the MS/CP operating frequency are constantly checked to make sure that there has been no failure in the track, track wires, or MS/CP. The presence of ac interference on the rails can prevent an accurate measurement of these characteristics, making the MS/CP think that there is a problem. This will cause the MS/CP to enter a right-side failure mode (see Chapter 5, Abnormal Operation of Railroad Equipment), and put the crossings warning devices into operation. The MS/CP will remain in this mode until some time after the measurements of the operating frequency come back within bounds. As in the case above, a timer keeps the warning devices activated for several seconds after things return to normal, to prevent intermittent interference from making the crossing go into and out of activation in a rapid fashion. 3. Damage to the Signal Equipment. In severe cases, ac interference on the rails will be of sufficient magnitude to do permanent physical damage to the circuits of the MS/CP. This usually takes the form of burned up circuit traces, capacitors, and transformers, but almost any type of damage is possible if the
9

This is often represented by the EZ number, Z1 or Z2 voltage, ED Voltage, RX number, or RX voltage, depending on equipment brand and vintage. (Always consult the equipment manufacturers documentation and/or technical support department.) This is often represented by the EX number, EDX Voltage, or Phase, depending on equipment brand and vintage. (Always consult the equipment manufacturers documentation and/or technical support department.)

10

9-29

Investigation

ac interference is high enough in voltage and current. Equipment has even been known to catch fire. If this happens, the most likely result will be a continuous activation of the grade crossings warning devices, but if the damage is extensive enough, the crossing lights may remain dark, and the bells silent. Fortunately, crossings equipped with gate mechanisms benefit from fail-safe design. In the event of a complete power failure to the gate mechanism, or a failure of the control voltage coming from the MS/CP or other controlling devices at the crossing, the arms will fall to their horizontal position by gravity alone. 4. Shunt Modulation. The presence of high levels (a few volts, or a few tens of volts rms) ac interference (usually 50/60 Hz) from one railroad rail to the other can detune some narrowband termination shunts. The termination shunts used to define the distant ends of the approaches to the grade crossing can be constructed in several ways. Most commonly, they consist of a series-resonant inductor-capacitor combination, which forms a short between the rails at a particular 11 frequency . This is called a narrowband termination shunt. Termination shunts may also be made of a length of #6 AWG (or heavier) wire, known as a hardwire termination shunt, or a large (>100,000 uF) capacitor, known as a wideband termination shunt. The problem with narrowband shunts is that they are not perfectly frequency-selective, and can pass a reduced amount of frequencies other than the one that they have been tuned for. If the frequency of the interfering ac energy is close enough to the tuned frequency of the narrowband termination shunt, large interfering ac currents may flow through it. If the amount of ac interference current passing through the termination shunt is sufficient to cause a significant degree of saturation of the magnetic core of the termination shunts inductor, the inductor will exhibit less inductance, and the shunt will become de-tuned. This re-tunes the series-resonant LC network to resonate at a new, higher frequency. If the amount of detuning was sufficient to move the center frequency of the shunt so that it coincided with the frequency being used by another audio-frequency track circuit (AFO, motion sensor, crossing predictor, etc.), the frequency-shifted termination shunt could, in theory, shunt the track effectively at that new resonant frequency. This could potentially shunt enough current away from the receiver of an audio-frequency overlay track circuit to cause the AFO to declare the track to be occupied by a train. And if this action proceeded at the proper rate with respect to time, it could also fool a MS/CP into believing that there was a train approaching. Furthermore, the now-detuned narrowband shunt would no longer be an effective shunt at its originally intended operating frequency. This could make the MS/CP using this narrowband termination shunt think that there was a problem with the termination shunt, the track, or the track wiring. This would result in an error condition being declared (see method 2, above), and rightfully so. 5. Other Track-Connected Device Modulation. Narrowband termination shunts are not the only devices that can be modulated by ac interference on the rails. The coupling networks of many types of track circuits can also be
11

Such narrowband termination shunts are actually tuned slightly low in frequency, in order to avoid a phenomenon known as hump-over, which can cause Motion Sensors to re-start a crossing shortly after a train leaves the far end of an approach.

9-30

Investigation

affected in a similar manner. As before, the interfering ac voltage must be present differentially, i.e. between the rails of the track. But the specific mechanism of this form of ac interference comes in two flavors: a. The coupling network for a track circuit gets de-tuned in a fashion that interferes with its normal operation. This can happen to the receiver and/or transmitter track-coupling networks of Motion Sensors, Crossing Predictors, and other audio-frequency track circuits. By saturating the receiver or transmitters track-coupling transformer, the transformers ability to pass the operating frequency of the track circuit is reduced. This reduction in signal amplitude can look like an incoming train. Again, Motion Sensors are more susceptible to this form of ac interference, because of their less-stringent requirements for activating the crossings warning devices. (See method 1 above) b. The track-coupling network of a track circuit can also be detuned, and act like a detuned narrowband termination shunt. If the de-tuning resulting from ac interference on the track causes it to effectively shunt the track and affect the operation of other track circuits, then the track circuit whose coupling network is detuned, and possibly other track circuits, can be affected.
Case #2: The Crossing Equipment is Activated Before it should be, but there is a Train Approaching

This is known as an early start, long warning time, or pre-ring12, depending on the exact timing of the symptoms, and can happen as a result of four of the five methods described above (1, 2, 4, and 5). What is different here is that the ac interference is usually transient, and caused by a connection made by the axles of the train as it enters the section of track where the affected track circuit is located. For mechanical reasons, insulated rail joints (insulated joints) are normally not installed in both rails of a track exactly opposite one another. This is because an insulated joint is slightly less stiff than the surrounding rails. Placing two insulated joints in a track directly across from one another creates a weak spot in the track, which subjects it to excessive flexing and rail movement. To avoid this, insulated joints are installed with a few feet of offset or stagger between them. This is usually only a few feet, but can be more than 10 feet in some cases. While the train is entering a section of track that is electrically isolated by insulated joints, there will come a time when the first axle of the train has entered the stagger, but has not yet passed completely into the section of track ahead (see figure below). This effectively connects one rail of the section of track ahead of the train to both rails of the section of track already occupied by the train. The unbalance created by this temporary connection between sections of track can briefly result in a greater conversion efficiency of induced ac voltages from common-mode to differential mode (see above). This means that the level of ac interference on rails in the section of track
12

A pre-ring is said to have occurred when the crossing is activated earlier than it should have been, recovers, and then is activated again at the correct time. This is sometimes the case with very slow trains. An early start or long warning time is said to have occurred if the crossing is activated much earlier than it should have been, and it remains activated until the entire train has passed the crossing.

9-31

Investigation

ahead of the train may rise suddenly when the first axle(s) of the train are in the stagger. This increased level of ac interference can adversely affect the operation of the Motion Sensor or Crossing Predictor at a crossing, and causing it to activate the warning devices. The duration of this transient event can be minimized by minimizing the amount of stagger between the insulated joints. But the duration will still remain a function of both the length of the stagger, and the speed of the train as it enters the stagger. The magnitude of the ac interference can be minimized by the careful application of narrowband shunts tuned for the interfering ac frequency. These serve to reduce the rail-to-rail voltage at the interfering ac frequency, and may be permissible as long as they do not interfere with the normal functioning of any nearby MS/CP. (180 Hz shunts installed across the track can potentially interfere with MS/CP units operating at adjacent frequencies such as 151, 156, 172, 210, or 211 Hz.) This set of symptoms can also be caused by incorrect programming or wiring of a MS/CP . Even simpler causes such as loose track wires or other electrical connections near the end of the section of track that the train is entering, or faulty insulator pads on steel or defective concrete railroad ties can cause this. The key to discriminating between such ordinary causes of pre-rings and early starts and ac interference related causes, is to look for a sudden rise in ac interference when a train enters the stagger. This is best done with an audio spectrum analyzer or frequency-selective voltmeter connected to the track. It may also be easier to test if two hardwire test shunts are used to simulate the presence of a train within the stagger. The procedure is simple: 1. Take all action necessary to protect the safety of train movements. 2. Measure the normal level of ac interference present on the track with no trains in the area. 3. Place one hardwire test shunt across the adjacent section of track that the train is leaving. 4. Place the second hardwire test shunt across the track within the stagger. 5. Measure the level of ac interference present 6. Remove both test shunts. If no significant rise in interference is seen with the head of a real or simulated train occupying the stagger and the adjacent section of track, then ac interference is an unlikely cause for this problem, and you should look for other causes, as listed above.
13

13

Some Motion Sensors and Crossing Predictors must be programmed via software or external programming jumpers on the hardware to expect sudden changes in track impedance. If this programming has not been done correctly, then the sudden appearance of a train within (or just outside of) an approach can cause the MS/CP to activate the warning devices at the crossing much earlier than it should.

9-32

Investigation

Case #3: The Crossing Equipment Stays Activated for Longer than it should after a Train has Passed Completely through the Grade Crossing

This is rarely a result of ac interference. When a train is occupying a section of track, the axles of the train effectively short the two rails together. With no appreciable voltage of any kind present between the rails, ac interference ordinarily cannot affect the operation of the Motion Sensor or 14 Crossing Predictor . Most often, an over-ring such as this is caused by the occurrence of an error condition (right-side failure mode) at some time during the trains approach to, or occupancy of, the grade crossing. Some of these right-side failure modes cannot be cleared until the train has moved far enough away from the grade crossing, and the required timers have expired. Some MS/CPs must see an indication that the tail of the train is indeed moving away from the crossing, before they will allow the crossing to recover. Some MS/CPs also have an unusually sophisticated island circuit built into them, which can detect the presence of poor shunting within the immediate vicinity of the crossing. Poor shunting can cause the island circuit to lose the train, and cause the crossing to recover prematurely, even if the train is still occupying the crossing. In order to avoid such incidents, these smarter island cards have timers in them that can delay the recovery of the crossing if evidence of poor shunting has been seen during a train move.
Case #4: The Crossing Equipment is Activated Briefly after a Train has Passed Completely Through the Grade Crossing and the Crossing has Recovered

This is known as a tail-ring. It can result from any of the same ac-interference-related and non-ac-interference-related reasons as the pre-ring or early start discussed in Case #2, above. Only the sequence of events is rearranged. Normally, only a single tail ring will occur, but multiple tail-rings are possible in some cases (e.g. multiple sets of insulated joints within an approach, and a slow train). From the ac interference perspective, tail-rings and pre-rings have identical causes. However, if the cause of the tail-ring is not ac interference, there is one additional cause that we should consider, that of hump-over. Hump-over is a non-linearity in the indicated position and/or direction of a train that can fool Motion Sensors and Crossing Predictors into believing that a train is approaching the crossing, even when it is really moving away from the crossing. Motion Sensors are particularly susceptible to this, as they may require very little apparent inbound motion to cause them to

14

This assumes that the train is shunting the track properly. Track circuits in the United States must be adjusted to respond to a shunt between the rails of 0.06 Ohms or less. Of course, they may be adjusted to respond to shunt resistances much higher than this. This is frequently done where infrequent train traffic, light axle loads, or rapid oxidation or contamination of the rail surfaces prevents the axles and wheels of a train from establishing a lowimpedance connection between the rails. This is known as poor shunting, and is a longstanding problem in some places. Ironically, the presence of ac interference can sometimes actually improve the operation of Motion Sensors and Crossing predictors by providing additional voltage to bite through the insulating and semiconductive films which can form on the rails.

9-33

Investigation

activate the crossings warning devices. Crossing Predictors have generally shown themselves to be immune to the effects of hump-over, but they are not necessarily so in all cases. Hump-over results from the imperfections of narrowband termination shunts. A perfect narrowband termination shunt would have zero impedance at the operating frequency of the Motion Sensor or Crossing Predictor. This is unattainable, due to the small but very real resistance of the wires and components (primarily the inductor) used to build the shunt. As a result of this resistance, after the last axle of a train passed a narrowband termination shunt, tuned for resonance at exactly the operating frequency of the MS/CP, as the train was on its way outbound from the crossing, the track impedance seen by the MS/CP would actually begin to drop again, after reaching a maximum value when the last axle was very close to (but beyond) the termination shunt. In the early days of Motion Sensors and Crossing Predictors, it was found that this problem could be alleviated by tuning the narrowband termination shunts for series-resonance at a slightly lower frequency than the operating frequency of the MS/CP. In this way, narrowband termination shunts are slightly different from tuned or tunable joint couplers (TJCs), which are tuned exactly on-frequency. There are generally two ways in which hump-over can come into play in causing tailrings. Either the existing narrowband termination shunt has become damaged or has drifted upwards in resonant frequency due to component aging, or a desperate signal maintainer has temporarily substituted a tunable joint coupler for a damaged narrowband termination shunt. In either case, the solution is the same, i.e. replace the narrowband termination shunt with a new or known good unit. As we can see, there are a number of ways in which ac interference can manifest itself in the behavior of Motion Sensors and Crossing Predictors. But few if any of these symptoms are sufficient conclusive proof of ac interference.

9-34

10
DIAGNOSTIC FLOW CHART
This chapter provides flow chart to facilitate root-cause analysis of potential ac interference conditions. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA.

10-1

Diagnostic Flow Chart

Introduction
The diagnostic flow chart provided in this chapter is intended to assist the reader in tracking down the root causes of ac interference related problems in railroad systems. In line with the approach to investigation presented in Chapter 9, the flow chart will take the user from the initial situation of A problem exists ac interference is suspected to one of four interim conclusions: Excessive common mode voltage Equipment damage from power system events Steady-state effects to equipment operation Problem is not caused by ac interference

From the interim conclusions above, the user is taken to a list of actions or mitigation that might be useful in addressing the problem. The user then evaluates the relative merits of various mitigation options to address the problem. Because site specific conditions vary, and because different railroads and power companies have a wide range of equipment and operating practices, more specific conclusions that are universally applicable can not be provided. The case studies in Chapter 14 are arranged in the same manner as the flow chart. The diagnostic flow chart is a process that will help narrow down the search for the root cause of the problem. It will work most of the time, but not all the time. If you are following the flow chart and things just dont make sense, then you may well be dealing with a case that the flow chart does not address effectively. When this occurs, please consider contributing your situation as a case study for future editions of the handbook. Also, any suggestions for improvements to the diagnostic flow chart, or any other aspect of the handbook, are always welcome.

10-2

Diagnostic Flow Chart

Following the Flow Chart


Where to begin To begin, start at the top of the next page with:
A problem exists ac interference is suspected

Follow the flow chart until you reach:

Go to A

or

Go to B

or

Go to C

And, go to the page beginning with: accordingly.

or

or

How to Proceed Diamond shaped boxes ask you to made a decision. These decisions will be based on observations and measurements. Sometimes the answer will be ambiguous. If that happens, consider both paths. Rectangular boxes usually ask you to do something. Sometimes you will take a measurement. Sometimes you will be directed to evaluate mitigation options. Most paths end with a rectangular box. If you get to the end of a path and your problems are not addressed, then look back through the flow chart again. You may have had more than one significant problem that needs to be addressed. Example: 60 Hz magnetic induction from a transmission line caused problems for signaling equipment. When the transmission line was de-energized for testing, the grade crossing equipment still had problems. It was determined that 180 Hz earth conduction from distribution facilities was effecting the crossing equipment. Two separate sources via two different paths needed to be addressed. Note: For 60 Hz power, the fundamental frequency is 60 Hz and the third harmonic is 180 Hz. For 50 Hz power, the fundamental frequency is 50 Hz and the third harmonic is 150 Hz.

10-3

Diagnostic Flow Chart


A problem exists ac interference is suspected Potential safety hazard consider special work rules such as working with insulated gloves Measure ac voltage rail-to-ground and across IJs (Amplitude and Frequency): Tests #1 and #2 If problem comes and goes (intermittent), try to make measurements while problem exists.

Does continuous ac voltage at any location exceed 50 Volts? No

Yes

Safety Hazard consider special work rules such as working with insulated gloves

Go to A

Go to C

No

Does problem involve damage to signal equipment?

Yes

Ask power company for a record of events such as faults, trips, and switching operations on all transmission and distribution lines near the affected track

No

Does the time of the damage correspond to the times of power system events?

Yes

Go to B

No

Did damage occur during an electrical storm?

Yes

Probably a direct lightning strike Work with power company to try to identify possible sources of power surges not previously reported. Sources could be power, railroad, or third party.

Figure 10-1 Railroad AC Power Interference Flow Chart

10-4

Diagnostic Flow Chart

Excessive Common Mode Voltage

A
Safety Hazard consider special work rules such as working with insulated gloves

Measured dominant frequency? 180Hz (Test #2) (50Hz) (150Hz) 60Hz Does a transmission line parallel the track? Yes Source is probably Transmission. Work with power company to identify any changes in equipment or operations. Consider changes to power, railroad, and third party facilities. Test insulated joints closest to the high voltage and one block to either side. A failed IJ will cause elevated voltages at adjacent IJs.

Source is Distribution, not Transmission

No

60Hz (50Hz)

For distribution to be the source of voltages this high the path is almost always conduction. Use Test #12 - Local Ground / Power Company Neutral Isolation Test to locate the path and source of the ac interference and Test #11 - Hardwire Shunt Testing

Eliminate the conductive path.

Look for inadvertent grounds. Grounding one end of a block candouble the induced voltage at the other end. -of-way to determine: If problem persists do a computer model of the right (1) are the voltages consistent with know factors, and (2) whatmitigation would be effective.

Power system mitigation to consider includes: Optimize phasing arrangements Minimize phase spacing Reduce load current Aerial or underground counterpoise Increase structure height Optimize phase current balance Split phasing Figure 10-1 (Continued) Railroad AC Power Interference Flow Chart

Railroad mitigation to consider includes: Shorten track blocks Eliminate by-pass couplers Underground counterpoise

10-5

Diagnostic Flow Chart

Equipment Damage from Power System Events

Potential safety hazard consider special work rules such as working with insulated gloves Remember that follow-on current can cause surge arrestor damage

Eliminate any connections between the railroads grounds and the power system grounds

Do a computer model of the right-of-way to determine: (1) Is the damage consistent with known factors, and (2) What mitigation would be effective.

Evaluate mitigation options

Power system mitigation involves reducing frequency of events, duration of events, and amplitude of current

Railroad mitigation involves increasing immunity to surges, reducing received current amplitude, and isolating equipment from surge currents

Mitigation to consider includes: Reduce frequency of events System modifications Reduce duration of events Limit reclosings Reduce fault clearing times Reduce amplitude of current Limit fault current Aerial or underground counterpoise

Mitigation to consider includes: Increase immunity to surges Use heavy duty arrestors Insure enclosure wiring minimizes exposure to induction Reduce received current amplitude Shorten track blocks Underground counterpoise Isolate equipment from surge currents Fuse track circuit leads

Figure 10-1 (Continued) Railroad AC Power Interference Flow Chart

10-6

Diagnostic Flow Chart

Steady-State Effects to Equipment Operation

C
Measure ac voltage rail-to-rail: Test #3

No

Is rail-to-rail voltage greater than 10% of rail-to-ground voltage?

Yes

Check power system equipment: Capacitor banks for blown fuses and failed capacitors Distribution neutral integrity Measure rail-to-rail frequency spectrum

Check track circuit balance: Test and repair IJs Test and repair arrestors (SPDs) Check by-pass couplers and termination shunts Eliminate conductive paths to ground and foreign sources

Is ac interference spectrum incompatible with the signal equipment present Yes

No

Concentrate on non-ac interference causes of the problem

Evaluate mitigation options

Power system mitigation to consider includes: Optimize phasing arrangements Minimize phase spacing Reduce load current Aerial or underground counterpoise Increase structure height Optimize phase current balance Split phasing Five-wire distribution

Railroad mitigation to consider includes: Increase equipment power output Adjust operating frequencies Shorten track blocks Improve track circuit balance Replace equipment for more immunity Underground counterpoise Eliminate IJs from crossing approach circuits Use heavy duty termination shunts

Figure 10-1 (Continued) Railroad AC Power Interference Flow Chart

10-7

11
GRAPHIC EVALUATION OF PROPOSED CHANGES BASED ON SIMULATIONS
This chapter provides a simple means for estimating the voltage and current induced onto a rail system by a power line for common important conditions.. Mr. Frazier has over 30 years of experience in the management and conduct of applied research programs and the application of the research results to the solution of system-specific problems. His major fields of experience include electromagnetic effects, electromagnetic compatibility, cathodic-protection systems, and detection systems. He is the author of the EPRI computer program CORRIDOR that has been extensively used by the engineering community to predict power-line interference on co-located structures. He is nationally known for his work on the electromagnetic compatibility of electric power, pipeline, and railroad systems. As President of Corr Comp Co., Mr. Frazier provides services to the power/pipeline/railroad industries in the pursuit of compatible common utility corridors.

11-1

Graphic Evaluation of Proposed Changes Based on Simulations

Introduction
The purpose of this chapter is to provide a simple means for estimating the voltage or current induced onto a rail system by a power line for common important conditions. The chapter will be of most use in making a preliminary evaluation of whether or not there appears to be a need for a more detailed study or analysis of an exposure, in preparation for an anticipated change in the parameters. Ideally, this information will serve as a tool in the initial routing/planning process to evaluate new transmission line routing and design options. Most often the impending change will be a new planned power line. However, other changes to the corridor may also be of interest, such as adding a new grade-crossing detector to an existing exposure, or changing the length of an existing track circuit. The chapter addresses five important compatibility issues: 1. Personnel safety for steady state magnetic field induction to rails. 2. Signal system equipment compatibility for steady state magnetic field induction to rails. 3. Personnel safety for steady state electric induction to open wire signal pole lines. 4. Personnel safety for faulted power-line magnetic field induction to rails. 5. Power-line fault induced rail current for lightning arrester survival. For each of these five issues a set of graphs has been prepared that will provide a near worst-case evaluation of the parameter of interest. There are many parameters that influence the coupling of energy from the power system to the rail system. In an effort to keep the number of graphs to a manageable and useful set, we have limited the number of parameters that are variable within the graphs. For parameters that are not variable within the graph set, we have attempted to set those parameters to a value that tends to produce worst-case conditions. For example, track ballast resistance can vary with seasonal and climatic conditions. For some evaluations, high ballast resistance conditions tend to produce the worst results. For other evaluations, low ballast resistance tends to produce the worst results. Rather than making the ballast resistance a variable in the graph set, where appropriate, we have set the ballast resistance to be a high or low value as appropriate to produce worst-case results. While the material of this chapter is intended to be self-explanatory on a stand-alone basis, review of Chapter 12 may provide additional insight into the effect of certain parameters on the results.

Summary of Graphical Aids for Compatibility Issues


A set of graphs is provided for each of the compatibility issues that are outlined above. One of the important parameters is the type of power line configuration under consideration. We have developed the material for this chapter using three basic power-line configurations. The three configurations are illustrated in Figure 11-1 (the vertical configuration), Figure 11-2 (the horizontal configuration), and Figure 11-3 (the horizontal delta configuration). Each figure shows the basic geometry that has been assumed to develop the curve set in this chapter. The 11-2

Graphic Evaluation of Proposed Changes Based on Simulations

spacing between power-line conductors is an important parameter in determining the fields from the line, and the spacing is generally related to the voltage class of the line. For many of the graphs of this chapter, the information is presented as a function of the conductor spacing. However, most of the time the key information that is available is the voltage class of the power line. Therefore we have shown the voltage class on the figures, where the spacing associated with each voltage class used for the development of the graphs is given in each of the figures (Figure 11-1 through Figure 11-3). We have only included single-circuit configurations in the graphs of this chapter. Double circuit lines have not explicitly been considered, since for most phase arrangements of the double circuit line, the worst condition is with one circuit deenergized. When one circuit is de-energized, the single-circuit arrangement is a good approximation. Each of these three figures shows a positive (+) and negative (-) side. Because of the conductor arrangement and shield wires, the field is sometimes different on one side of the line than the other (asymmetric). For the sake of clarity, the positive and negative sides are shown in each figure with the power flow in all cases being into the page. The positive and negative side references are also shown in the tables which direct the reader to the proper figures. Small unbalances in the phase currents can have a significant effect on the value of induced voltage. Therefore the effect of unbalanced current is included in the graphs. We have used the measured unbalanced currents of Reference [4] in the development of these graphs. For use in this material, the measured data of [4] have been converted to an approximate equivalent residual current. Figure 11-4 shows the result of that conversion, and is the residual current that has been assumed for developing the curves of this chapter. There is no good data available on steady state unbalance current. We expect that most lines (particularly transmission lines) will have less current unbalance than assumed for these curves. Therefore, these curves should tend to be worst case.

11-3

Graphic Evaluation of Proposed Changes Based on Simulations

Figure 11-1 Vertical Power Line Configuration

11-4

Graphic Evaluation of Proposed Changes Based on Simulations

Figure 11-2 Horizontal Power Line Configuration

11-5

Graphic Evaluation of Proposed Changes Based on Simulations

Figure 11-3 Delta Power Line Configuration

The procedure used to include the effects of the unbalance current has been explained in Chapter 12. The balanced contribution and the unbalanced contribution caused by the power-line currents have been separately determined and the results have been algebraically added, to produce an in-phase result, which is worst case. The following subsections outline the use of the graphs for each of the compatibility issues that have been outlined above. 11-6

Graphic Evaluation of Proposed Changes Based on Simulations

Figure 11-4 Residual Current from Cramer Data

Personnel Safety for Steady State Magnetic Field Induction to Rails The purpose of this set of graphs is to determine the maximum average phase current that will result in 25 volts rail-to-ground at a rail insulated joint for a given power-line voltage, line configuration, soil resistivity, and track-to-power line separation. It is generally desired to maintain the steady-state induced voltage to a value that is less than 25 volts. Therefore, the average phase current should be less than the value that is shown in the figures, for the other conditions of interest. All of the curves are for a track ballast resistance of 100 ohmkft (commonly referred to as 100 ohms per 1000 feet by railroads), which is a high ballast resistance condition and a near worstcase for induced rail-to-ground voltage. All the curves are for a one mile long track circuit. To obtain the average phase current to give 25 volts rail-to-ground for other lengths, divide the current value obtained for the one-mile track circuit by the length of the track circuit of interest, in miles. Thus, to obtain the power line current that will induce 25 volts onto a track circuit that is 2 miles long, divide the current shown in the appropriate graph by 2. Table 11-1 provides a guide to determining the figure number that corresponds to the desired conditions. The table shows that two alternatives for overhead shield wires have been provided. The two conditions are: segmented shield wires or no shield wires multigrounded 7 # 9Alumoweld shield wires

The results for steel overhead shield wire (OHSW) will lie between these two cases and has not been provided as separate figures, except where noted.

11-7

Graphic Evaluation of Proposed Changes Based on Simulations

The curves for steady state personnel safety are all labeled as Figure 11-5 through 11-40. A seemingly odd characteristic occurs in Figures 11-29 through 11-40, which are the curves for the horizontal delta configuration. All of these curves show a significant decrease in allowable current for a 500-kV line relative to for a 345-kV line which is a result of the conductor 1 separations for the different voltage classes.
Table 11-1 Personnel Safety for Steady State Magnetic Field Induction to Rails Figure Number Shield Wire Earth Resistivity ohmm

Circuit Orientation Horizontal Delta

Side

Vertical 11-5 11-6 11-7 11-8 11-9 11-10 11-11 11-12 11-13 11-14 11-15 11-16 11-17 11-18 11-19 X X X X X X X X X X X X

Horizontal

Negative X

Positive 10 X 10 100 X 100 1000 X 1000 Shield Wire X Shield Wire Shield Wire X Shield Wire Shield Wire X Shield Wire 10 10 100 100 1000 1000 10 X 10 100

X X X

The horizontal and vertical conductor spacing for the horizontal delta configuration are about the same for each line type except for the 500-kV line. The 500-kV line horizontal conductor spacing is greater than the vertical spacing. This means that the phase conductors are spaced much farther apart, which provides less cancellation of the fields than the other voltage classes of this configuration.

11-8

Graphic Evaluation of Proposed Changes Based on Simulations

Table 11-1 (Continued) Personnel Safety for Steady State Magnetic Field Induction to Rails Figure Number Shield Wire Earth Resistivity ohmm

Circuit Orientation Horizontal Delta

Side

Vertical 11-20 11-21 11-22 11-23 11-24 11-25 11-26 11-27 11-28 11-29 11-30 11-31 11-32 11-33 11-34 11-35 11-36 11-37 11-38 11-39 11-40

Horizontal X X X X X X X X X

Negative

Positive X 100 1000 X 1000 Shield Wire X Shield Wire Shield Wire X Shield Wire Shield Wire X Shield Wire 10 10 100 100 1000 1000 10 X 10 100 X 100 1000 X 1000 Shield Wire X Shield Wire Shield Wire X Shield Wire Shield Wire X Shield Wire 10 10 100 100 1000 1000

X X X X X X X X X X X X

11-9

Graphic Evaluation of Proposed Changes Based on Simulations

Signal System Equipment Compatibility for Steady State Magnetic Field Induction to Rails The purpose of this set of graphs is to determine the average phase current that will result in 5 volts rail-to-rail for a given power-line voltage, line configuration, soil resistivity, and track-topower line separation. The rail-to-rail voltage is caused by a shorted rail insulated joint (IJ) and the analysis assumes a train shunt at an adjacent IJ that bridges across that IJ. The shorted IJ and occupied condition tends to be near worst case for causing rail-to-rail voltage for a given excitation by power line magnetic-field coupling. The track ballast assumed for this set of graphs is 100 ohmkft, which is also near worst case. The voltage that is evaluated in the analysis is the rail-to-rail voltage at the shorted IJ location, on the opposite side of the IJ from the train shunt. That voltage tends to be the highest voltage that results from the shorted IJ and train shunt; although, for some conditions the highest voltage is observed at the adjacent IJ to the shorted one that does not have the train shunt. However, the voltage that is shown in these graphs, at the shorted IJ, is within a few percent of the highest voltage that is observed. The 5-volt condition that has been assumed for this set of graphs is based on the AREMA 5-volt susceptibility guideline for highway crossing equipment. Not all highway crossing equipment is tolerant to 5 volts of 60 Hz. If less tolerant units are to be used, the transmission current can be directly scaled by the ratio of the susceptibility voltage to 5 volts. The rail-to-rail voltage for the shorted IJ condition does not scale linearly with track-circuit length. Therefore, this graph set is provided for two track-circuit lengths, namely one and two miles long. Intermediate track circuit lengths can be interpolated. Table 11-2 provides a guide to determining the figure number that corresponds to the desired conditions. The curves for steady state equipment compatibility are all labeled as Figure 11-41 through 11-112. An irregularity occurs in Figures 11-65 through 11-76 and Figures 11-101 through 11-112, which are the curves for the horizontal delta configuration. All of these curves a show a decrease in allowable current for a 500-kV line from that of a 345-kV line.

11-10

Graphic Evaluation of Proposed Changes Based on Simulations

Table 11-2 Signal System Equipment Compatibility for Steady State Magnetic Field Induction to Rails Figure Number Circuit Orientation Vertical 11-41 11-42 11-43 11-44 11-45 11-46 11-47 11-48 11-49 11-50 11-51 11-52 11-53 11-54 11-55 11-56 11-57 11-58 11-59 11-60 11-61 11-62 11-63 11-64 11-65 11-66 11-67 11-68 11-69 11-70 11-71 11-72 11-73 11-74 11-75 11-76 11-77 X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Horizontal Horizontal Delta Side Negative X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Positive 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 Shield Wire Earth Resistivity ohmm Track Block Length in Miles 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 2

11-11

Graphic Evaluation of Proposed Changes Based on Simulations

Table 11-2 (Continued) Signal System Equipment Compatibility for Steady State Magnetic Field Induction to Rails Figure Number Circuit Orientation Vertical 11-78 11-79 11-80 11-81 11-82 11-83 11-84 11-85 11-86 11-87 11-88 11-89 11-90 11-91 11-92 11-93 11-94 11-95 11-96 11-97 11-98 11-99 11-100 11-101 11-102 11-103 11-104 11-105 11-106 11-107 11-108 11-109 11-110 11-111 11-112 X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Horizontal Horizontal Delta Side Negative X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Positive X 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 10 10 100 100 1000 1000 Shield Wire Earth Resistivity ohmm Track Block Length in Miles 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2

11-12

Graphic Evaluation of Proposed Changes Based on Simulations

Personnel Safety for Steady State Electric Induction to Open Wire Signal Pole Lines The purpose of this set of graphs is to determine the current that will flow through a person who touches an open-wire signal pole line conductor, excited by electric induction, for a given powerline voltage, line configuration, and track-to-power line separation. The curves are presented as the person current in milliamperes per mile of conductor. The allowable current depends on the criterion that is used for assessing safety. Chapter 7 provides discussion of the considerations. Representative value of safe let-go body current for males is approximately 10-mA. Knowledge of the signal circuit lengths that are exposed and this set of curves will allow the body current to be evaluated for a given power-line voltage and separation. The body current is the value shown on the curve set times the length of the open-wire circuit in miles. Table 11-3 provides a guide to determining the figure number that corresponds to the desired conditions. The curves for steady state electric induction to open wire signal pole lines are all labeled as Figure 11-113 through 11-124. An irregularity occurs in Figures 11-121 through 11-124, which are the curves for the horizontal delta configuration. All of these curves show a sharp increase in current through the person for a 500-kV line from that of a 345-kV line where a a less drastic increase would otherwise be expected.
Table 11-3 Personnel Safety for Steady State Electric Induction to Open Wire Signal Pole Lines Figure Number Vertical 11-113 11-114 11-115 11-116 11-117 11-118 11-119 11-120 11-121 11-122 11-123 11-124 X X X X X X X X X X X X Circuit Orientation Horizontal Horizontal Delta Negative X X X X X X X X X X X X Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Shield Wire Side Positive Shield Wire

Personnel Safety for Faulted Power-Line Magnetic Field Induction to Rails The purpose of this set of graphs is to determine the maximum faulted phase to ground current that will result in a safe rail-to-ground touch potential at the end of a long exposure for which the track arresters at IJ locations fire as a result of the induced rail voltage. These conditions tend to 11-13

Graphic Evaluation of Proposed Changes Based on Simulations

result in the highest rail-touch potential. The graphs have been prepared for assumed high track ballast resistivity (100 ohmkft). The allowable fault current is plotted as a function of the trackto-power line distance. Separate curves are given for different fault-clearing times that range from 3 cycles to 25 cycles. Graphs are presented for different values of body weight (110 or 155 lbs) and different values of soil resistivity. Table 11-4 provides a guide to determining the figure number that corresponds to the desired conditions. The curves for faulted power-line magnetic field induction to rails are all labeled as Figure 11-125 through 11-130.
Table 11-4 Personnel Safety for Faulted Power-line Magnetic Field Induction to Rails Figure Number 11-125 11-126 11-127 11-128 11-129 11-130 Weight of Person lb 155 110 155 110 155 110 Earth Resistivity ohmm 10 10 100 100 1000 1000

Power-Line Fault Induced Rail Current for Lightning Arrester Survival The purpose of this set of graphs is to determine the approximate phase-to-ground fault-induced rail current for a long uniformly-excited exposure. The induced rail current is directly related to the fault-current and is dependent on the distance between the track and the power line and the soil resistivity. If the fault-current tolerance of the arrester is known, the maximum allowable fault current can be determined by this set of curves. Of course the maximum tolerable arrester current is dependent on the duration of the fault. The maximum tolerable arrester current is also dependent on the specific design of the arrester. The graphs of this group assume a uniformly excited region of track that is sufficiently long to allow the current to approach the long approximation. The results are not very dependent on the specifics of the power configuration, but do depend on the type of OHSW and its location. For these curves, we have assumed a high-strength steel OHSW, which is a typical construction and does not provide significant shielding or reduction of the power line fault-current field. There is only one graph that is presented to relate the induced rail current to the fault current for a long exposure (Figure 11-131). The other graph in this group tries to quantify the question of what is a sufficiently long exposure to obtain the long approximation? Figure 11-132 shows for different soil resistivities how the induced track current deviates from the long approximation, versus the exposure length. Table 11-5 provides a guide to determining the figure number that corresponds to the desired conditions. The curves for faulted power induced rail current are labeled as Figure 11-131 and 11-132. 11-14

Graphic Evaluation of Proposed Changes Based on Simulations

The following is an example of a preliminary assessment of compatibility for a proposed transmission line and an existing rail system. The example illustrates how the Chapter 11 material has been used to assess the need for more detailed investigation.
Table 11-5 Power-Line Fault Induced Rail Current for Lightning Arrester Survival Figure Number 11-131 11-132 Exposure Type Long Varying Length

Example of Preliminary Compatibility Assessment by Graphical Procedures


Introduction and Background A new 115 kV electric transmission line and distribution substation adjacent to and parallel with an existing rail line is proposed. A preliminary review using graphical procedures has been performed to determine if: the planned exposure has no obvious electromagnetic impact on the rail system, for which additional investigation is not needed. the effect of the electromagnetic coupling on the rail system is uncertain, for which further investigation is needed to resolve. the expected electromagnetic coupling is likely cause personnel safety or signal-equipment compatibility issues which may require mitigation. Further investigation will be necessary.

The power company has provided information for use in making a preliminary review. Specific information on the proposed transmission system includes: The line voltage is 115 kV, single-circuit 3-phase. Initial line current of 40 amps. Ultimate line current 570 A (summer) and 759 A (winter). Single pole wood structures with 400-ft spacing. Phase conductors are 336 ACSR Linnet (26/7). Horizontal delta phase arrangement (one phase on the track side of the pole). Horizontal post insulators. Minimum vertical ground clearance along transmission centerline 24 ft. Centerline of transmission line is approximately 40 ft west of near track centerline. Transmission line located adjacent to tracks from MP 92.1 to 94.5. Maximum phase-to-ground fault current within exposure is 3000 amperes. Fault clearing time is 6 cycles (0.1 seconds). 11-15

Graphic Evaluation of Proposed Changes Based on Simulations

The transmission line terminates at a proposed substation on a four-acre site near the south end of the railroad parallel.

Relevant information on the track system includes: Two dc-circuit signaled tracks separated by approximately 15 feet. Open-wire signal pole line, approximately 20 feet east of tracks. Approximately two-mile long signal blocks. IJs and signals at MP 93, near mid-region of exposure. Signaled grade crossing (210-Hz HXP) near south end of exposure. Signaled grade crossings (114-Hz and 151-Hz HXPs) approximately -mile north of exposure. Narrow band shunts rail-to-rail at the crossing approach terminations within exposure at crossing frequencies.

The compatible operation of the railroad and power systems must consider both the steady state and faulted conditions of the power system. Concerns for the steady state operation of the power system include: electric-current shock safety for railroad maintenance personnel and the public, and compatible operation of the railroad signal equipment.

Concerns for the power line steady state compatibility with the railroad signaling systems include: public safety associated with the safe operation of Highway Grade Crossing signal and control equipment, and train delays and safe operation of train-control signal systems.

Concerns for the power-line fault conditions include: personnel safety associated with touch potentials at equipment housing locations, induced current that may exceed the ratings of signal-system lightning protectors, resulting in destruction of signal equipment, voltage stress on railroad signal equipment and cables.

Preliminary Review of Coupled Voltage and Current


Steady State Coupling to Track

The most prevalent cause of rail-system induced voltage is the current that flows in the power-system conductors. The current in the power-system conductors couples voltage onto the

11-16

Graphic Evaluation of Proposed Changes Based on Simulations

railroad system by magnetic-field induction, similar to a simple single-turn transformer. Two important power-system parameters that influence the induced rail-system voltage are: the spacing between the power conductors. the unbalance of the current in the three power phase conductors,

The unbalance of three-phase power-line currents can be an important influence on the voltage induced onto nearby railroad conductors. Transmission line current unbalance is influenced by the spacing of the power conductors to each other and to the earth, the loading of the line, and the power factor of the phases. Personnel shock safety becomes a concern for rail-to-earth induced steady state voltage in the range of 25 volts. The voltage induced onto rails that parallel a power transmission line is dependent on several parameters including, the transmission line current, the balance of the current, the transmission line configuration, the spacing between phase conductors, the distance between the rails and the power line, and the soil resistivity. The proposed 115-kV line has a horizontal delta phase arrangement with one phase conductor on the side of the track. Referring to Figure 11-3 for a delta configuration shows that the single phase conductor is on the positive side, such that the curves of Figure 11-32 (for 100-ohmm soil) may be representative. The distance to the two tracks from the transmission line is approximately 40 and 55 feet. The curves of Figure 11-32 show that for 50-ft separation from a representative 115-kV line, approximately 0.8 kA line current or higher, with unbalance as in Figure 11-4, can result in 25 volts or more on a uniformly excited 1.0-mile track circuit. Since the track circuits for this exposure are approximately 2 miles long, we must divide the current value obtained from Figure 11-32 by the ratio of (actual track circuit length)/(1 mile) = 2. Therefore, the graphical estimation procedure of this chapter indicates that personnel safety issues may occur if the transmission line current is higher than 0.4 kA, or 400 amps. The Inductive Coordination Information sheet provided by the power company indicates an initial line current of 40 amps, with ultimate 570 A (summer) and 759 A (winter). The ultimate currents are in a range to cause induced voltage in the range for personnel safety concern, for the other conditions assumed for the analysis. The track circuits in this region are approximately 2 miles long, but the location of the IJs do not result in uniform excitation of a complete track circuit. Because of the IJ arrangement in the proposed excitation region, the induced voltage may not exceed 25 volts, however, a more specific assessment would necessitate a detailed modeling to evaluate the specific geometry of the exposure. An important element of inductive coordination is to identify if existing railroad signal equipment is compatible with the induced rail-to-rail voltage or insulated joint voltage that may occur, even if personnel-safety guidelines are met. Tuned joint couplers can be susceptible to quite low induced voltage across the coupled insulated joints. The induced voltage between two rails is typically within a few percent of the induced rail-to-ground voltage for nominally balanced track conditions, but can be considerably higher for some normal unbalancing conditions. Many in the railroad industry consider a single degraded rail insulated joint (IJ) to be a normal operating condition. The degraded IJ is considered to be a normal condition because when no significant induced voltage is present, many signal systems will function properly with a single degraded joint. However, in the presence of significant induced rail-ground voltage, a 11-17

Graphic Evaluation of Proposed Changes Based on Simulations

shorted or degraded-resistance insulated joint can result in non-negligible rail-to-rail voltage that may interfere with the normal operation of the signal system. Important unbalancing conditions should be evaluated in a compatibility investigation and for mitigation design. Common unbalancing conditions may include an unintentionally grounded track lead, or a shorted insulated joint (with or without a train within the affected track region). The rail-to-rail induced 60 Hz voltage tolerance for several types of track signal systems are specifically identified in AREMA Manual Parts. Table 11-6 summarizes the AREMA target interference-tolerance voltages for several types of track signal equipment. The guideline tolerance for most grade-crossing equipment is 5 volts.
Table 11-6 AREMA Guidelines for Interference Tolerance of Selected Signal Equipment C&S Manual Part 3.1.20 3.1.26 8.2.1 Equipment Type Motion Sensitive Systems to Control Highway Grade Crossings Constant Warning Time Systems to Control Highway Grade Crossings Audio Frequency Track Circuits Interference Tolerance 5 V rms 60 Hz 5 V rms 60 Hz 10 V rms, 60-180 Hz

The information of Table 11-2 for signal system equipment compatibility for steady state magnetic field induction to rails can help locate the appropriate Chapter 11 figure to estimate the line current that may cause up to 5 volts rail-to-rail for unbalanced track. It is found that Figure 11-104 has relevant parameters for the present example: Horizontal delta power circuit arrangement. Positive distance from the transmission line to the track.
Does not have Alumoweld OHSW.

100-ohmm soil resistivity. Two mile long track circuit.

From Figure 11-104 it is found that a transmission line circuit with approximately 0.08 kA (80 amperes) may result in up to 5 volts rail to rail for a two mile long track circuit. The 5 volts will appear at one of the insulated joints, and necessitates exposure of approximately three contiguous track circuits, which will not occur for this example exposure. The specific requirements on induced rail-to-ground voltage and rail-to-rail voltage is dependent on the specific signal equipment used and the arrangement of the signal equipment within the track-signal circuit. The results are dependent on where the IJs are located within the exposure and where the grade-crossing equipment is located within the signal block. Site-specific signal designs, such as the use of wide-band shunts (WBS) at insulated joints can significantly alter the induced rail-to-rail voltage at grade-crossing sites. Some grade-crossing signal systems or components do not meet the above AREMA guidelines. Therefore where grade-crossing equipment is present, it may be prudent to evaluate the effect of possible track unbalancing 11-18

Graphic Evaluation of Proposed Changes Based on Simulations

conditions on the rail-to-rail voltage at the input to the signal equipment, if the expected transmission line current exceeds the value estimated by using the procedures of Chapter 11. Since the anticipated ultimate current for the proposed 115-kV transmission line exceeds the current determined by use of Figure 11-104, a more detailed site-specific compatibility study may be prudent.
Steady State Coupling to Pole Line

The existing track signal system is dc with associated signal pole line. Capacitive coupling (electric induction) from transmission lines can induce current and voltage onto long high impedance conductors, such as pole-line open wire conductors. A primary concern is for personnel safety with regard to the current that can flow through a person who contacts one of the energized conductors. Safety guidelines recommend limiting the available capacitive coupled current from 5 mA to 10 mA. The National Electric Safety Code, Rule 232.C.2.c recommends A safety threshold of 5 mA, recommended for the general population including children.... Whereas the ITU (CCITT) recommends a capacitive coupling current limit of 10 mA2. Table 11-3 can be used to help quantify the personnel safety concern for the existing signal pole line. From Table 11-3, we identify Figure 11-118 as appropriate for this example: Horizontal delta power circuit arrangement. Positive distance from the transmission line to the track.
Does not have Alumoweld OHSW.

Figure 11-118 shows the current collected by an insulated conductor per mile of parallel to a 115-kV vertical geometry transmission line with typical phase spacing, as a function of separation between the conductor and the transmission line. If the pole-line is on the opposite side of the railroad ROW as the proposed transmission line, it will be at least 65 ft from the transmission line. The curve of Figure 11-118 shows that at that separation, the expected induced current is approximately 3 mA per mile of exposure. For a pole-line conductor that runs the length of the exposure (approximately 2.5 miles) the expected induced current is approximately 7.5 mA, which is between the safety guideline of 5 mA or10 mA. Therefore, a more detailed investigation may be desirable to quantify the electric and magnetic field induction to the pole line.
Fault Current Coupling to Track

A single phase-to-earth fault along a transmission line results in a high current in the faulted phase, which can be considerably higher than the normal load current. The high current only exists for a brief period, since sensing circuits on the power system will open breakers to stop the

ITU, Directives concerning the protection of telecommunication lines against harmful effects from electric power and electrified railway lines, Volume VI, Danger and Disturbance, Geneva 1989, ISBN 92-61-04041-3.

11-19

Graphic Evaluation of Proposed Changes Based on Simulations

flow of current. The high fault current in the transmission line conductor creates a high magnetic field, which can induce high voltage and current in long parallel conductors such as a rail system. The principal concerns for the power-line fault conditions include: personnel safety associated with touch potentials along the tracks and at equipment housing locations, induced current that may exceed the ratings of signal-system lightning protectors, resulting in damage to signal equipment,

Table 11-4 can be used to help estimate the personnel safety concern for fault-induced rail voltage. From that table the appropriate figure for a 155-lb or 110-lb person can be selected. Figure 11-128 is appropriate for a 110-lb person. That figure group provides the maximum expected rail-to-ground voltage induced by magnetic field coupling from a single-phase to ground fault. The figure shows that at 40 feet from the transmission line and a 6-cycle fault, the allowable fault current to maintain a safe rail touch potential is approximately 800 amperes for a long exposure. The expected fault current is 3000 amps, but the exposure is not long, therefore, it is not evident that there will be no problem for personnel safety for fault conditions, without more detailed investigation. Conclusions The following conclusions are drawn from the preliminary review of the proposed 115-kV transmission line electromagnetic compatibility with the existing railroad signal system. The ultimate steady state transmission line currents are expected to cause induced rail voltage in the range for personnel safety concern (25 volts rail-to-ground). The induced voltage may not exceed 25 volts, because of the specific block arrangement in the proposed excitation region. However, a more specific assessment would necessitate a detailed modeling to evaluate the specific geometry of the exposure. To maintain rail-to-rail voltage that is less than the AREMA 5-volt susceptibility guidelines for grade crossing equipment may necessitate transmission line current that is less than 80 amperes. The initial proposed current for the transmission line is only 40 amperes. However, the ultimate current is expected to exceed 500 amperes. Assessment of induced voltage at track signal equipments for normal possible track unbalancing conditions will necessitate a site-specific compatibility study and may necessitate the development of mitigation measures. The existing track signal system is dc with associated signal pole-line. Personnel safety issues may be caused by the capacitive coupling to the open-wire pole-line, if the wire circuits extend the length of the exposure. A preliminary review of transmission line fault-current condition indicates that the induced rail voltage may exceed a safe touch potential for the anticipated fault current. A more detailed evaluation of the specific system geometry is necessary to evaluate the need for mitigation. Such an evaluation would also include assessment of induced rail current that may flow through fired lightning arresters, which was not addressed in this preliminary review.

11-20

Graphic Evaluation of Proposed Changes Based on Simulations


7

6 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side

500 ft Power Pole-to-Track Distance (ft) 35 40


35 Power Pole-to-Track Distance (ft) 40

5 300 ft 4

200 ft

150 ft 100 ft

50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 345 kV 25 Phase Separation (ft) 30 500 kV

Figure 11-5 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m

6 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side

500 ft

300 ft

4 200 ft 3 150 ft 2 100 ft 1 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

Figure 11-6 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m

11-21

Graphic Evaluation of Proposed Changes Based on Simulations

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side

800 ft

300 ft

200 ft 150 ft

100 ft 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-7 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m

A v e ra g e P h a s e C u rre n t M a g n itu d e (k A )

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft 5 Positive Side

800 ft

500 ft 4 300 ft 3 200 ft 2 150 ft 100 ft 1 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 345 kV Phase Separation (ft) 25 30 500 kV 35 40 P o w e r P o le -to -T ra c k D is ta n c e (ft)

Figure 11-8 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m

11-22

Power Pole-to-Track Distance (ft)

500 ft

Graphic Evaluation of Proposed Changes Based on Simulations


4 3.5 Average Phase Current Magnitude (kA) 3 2.5 300 ft 2 1.5 1 0.5 0 0 5 10 138 kV 15 230 kV 20 25 345 kV 34 kV 115 kV Phase Separation (ft) 30 500 kV 35 40 200 ft 150 ft 100 ft 50 ft

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side

800 ft

Figure 11-9 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m

4 3.5 Average Phase Current Magnitude (kA) 3 2.5 2

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side

800 ft

500 ft Power Pole-to-Track Distance (ft) 35 40

300 ft

200 ft 1.5 1 0.5 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 150 ft 100 ft 50 ft

Figure 11-10 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m

Power Pole-to-Track Distance (ft)

500 ft

11-23

Graphic Evaluation of Proposed Changes Based on Simulations

10 9 Average Phase Current Magnitude (kA) 8 7 6 5

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wire

500 ft

300 ft 4 3 200 ft 2 1 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 35 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 40


35 Power Pole-to-Track Distance (ft) 40

Figure 11-11 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m

12

Average Phase Current Magnitude (kA)

10

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wire

500 ft

6 300 ft 4 200 ft 2 150 ft 100 ft 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

Figure 11-12 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m

11-24

Graphic Evaluation of Proposed Changes Based on Simulations

5 4.5 Average Phase Current Magnitude (kA) 4 3.5 3 2.5

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wire

800 ft

500 ft

300 ft 2 1.5 1 0.5 0 0 5 34 kV 10 138 kV 15 230 kV 20 25 345 kV 115 kV Phase Separation (ft) 30 500 kV 35 200 ft 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 40 35 Power Pole-to-Track Distance (ft) 40

Figure 11-13 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m
5 4.5 Average Phase Current Magnitude (kA) 4 3.5 500 ft 3 2.5 2 1.5 1 0.5 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 300 ft 200 ft 150 ft 100 ft 50 ft

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wire

800 ft

Figure 11-14 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m

11-25

Graphic Evaluation of Proposed Changes Based on Simulations

2.5

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wire

800 ft

1.5

300 ft 200 ft

150 ft 100 ft

0.5

50 ft

0 0 5 34 kV 10 138 kV 15 230 kV 20 25 345 kV 115 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-15 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m

2.5

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wire

800 ft

1.5

300 ft 200 ft 150 ft

1 100 ft 50 ft 0.5

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-16 Maximum Average Line Current for 25V Rail-to-Ground, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m

11-26

Power Pole-to-Track Distance (ft)

500 ft

Power Pole-to-Track Distance (ft)

500 ft

Graphic Evaluation of Proposed Changes Based on Simulations

4.5 4 Average Phase Current Magnitude (kA) 3.5 3 2.5 2 1.5

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Power Pole-to-Track Distance (ft) 35 500 kV 40 35 500 kV Power Pole-to-Track Distance (ft) 40

500 ft

300 ft 200 ft

1 0.5 0 0 5 10 15 34 kV 115 kV 138 kV 230 kV 20 345 kV 25 Phase Separation (ft) 30

150 ft 100 ft 50 ft

Figure 11-17 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m

4.5 4 Average Phase Current Magnitude (kA) 3.5 3 2.5 2 1.5

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side

500 ft

300 ft 200 ft

1 0.5 0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

150 ft 100 ft 50 ft

Figure 11-18 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m

11-27

Graphic Evaluation of Proposed Changes Based on Simulations

4 3.5 Average Phase Current Magnitude (kA) 3 2.5

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side

800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40


35 500 kV Power Pole-to-Track Distance (ft) 40

500 ft 2 1.5 1 0.5 0 0 5 10 15 34 kV 115 kV 138 kV 230 kV 20 345 kV 25 Phase Separation (ft) 30 300 ft 200 ft 150 ft 100 ft 50 ft

Figure 11-19 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m

4 3.5 Average Phase Current Magnitude (kA) 3 2.5

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side

800 ft

500 ft 2 1.5 1 0.5 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 300 ft 200 ft 150 ft 100 ft 50 ft

Figure 11-20 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m

11-28

Graphic Evaluation of Proposed Changes Based on Simulations

Average Phase Current Magnitude (kA)

2.5

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side 800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40
35 500 kV Power Pole-to-Track Distance (ft) 40

2 500 ft 1.5 300 ft 1 200 ft 150 ft 100 ft 50 ft

0.5

0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

Figure 11-21 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m

Average Phase Current Magnitude (kA)

2.5

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side 800 ft

2 500 ft 1.5 300 ft 1 200 ft 150 ft 100 ft 50 ft

0.5

0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

Figure 11-22 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m

11-29

Graphic Evaluation of Proposed Changes Based on Simulations

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wires Power Pole-to-Track Distance (ft) 40 35 500 kV Power Pole-to-Track Distance (ft) 40

4 500 ft 3

300 ft 200 ft

150 ft 100 ft 50 ft

0 0 5 10 15 25 230 kV 20 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 35

Figure 11-23 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m

5 4.5 Average Phase Current Magnitude (kA) 4 3.5 3 2.5 2 1.5 1 0.5 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wires

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

30

Figure 11-24 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m

11-30

Graphic Evaluation of Proposed Changes Based on Simulations

4.5 4 Average Phase Current Magnitude (kA) 3.5 3 2.5 2

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wires

800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40

500 ft

300 ft 1.5 1 0.5 0 0 5 10 15 34 kV 115 kV 138 kV 230 kV 20 345 kV 25 Phase Separation (ft) 30 200 ft 150 ft 100 ft 50 ft

Figure 11-25 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m

4 3.5 Average Phase Current Magnitude (kA) 3 2.5 2 1.5 1 0.5 0 0

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wires Power Pole-to-Track Distance (ft) 35 500 kV 40 800 ft

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft)

30

Figure 11-26 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m

11-31

Graphic Evaluation of Proposed Changes Based on Simulations

3.5

3 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wires

800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40 35 500 kV Power Pole-to-Track Distance (ft) 40

2.5 500 ft 2

1.5

300 ft 200 ft

150 ft 100 ft 50 ft

0.5

0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

Figure 11-27 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m

2.5

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wires

800 ft

500 ft 1.5

300 ft 1 200 ft 150 ft 100 ft 50 ft

0.5

0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

Figure 11-28 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m

11-32

Graphic Evaluation of Proposed Changes Based on Simulations

6 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-29 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side

500 ft

3 300 ft 2 200 ft 150 ft 100 ft 50 ft

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-30 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m

11-33

Graphic Evaluation of Proposed Changes Based on Simulations

5 4.5 Average Phase Current Magnitude (kA) 4 3.5 3 2.5 2 1.5 1 0.5 0 0

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side

800 ft Power Pole-to-Track Distance (ft) 30

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft)

345 kV

25 500 kV

Figure 11-31 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m

4.5 4 Average Phase Current Magnitude (kA) 3.5 3

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side

800 ft

500 ft 2.5 2 1.5 1 0.5 0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 300 ft 200 ft 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 30

Figure 11-32 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m

11-34

Graphic Evaluation of Proposed Changes Based on Simulations

3.5

3 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side

800 ft Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

2.5 500 ft 2 300 ft 1.5 200 ft 1 150 ft 100 ft 50 ft 0.5

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-33 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m

Average Phase Current Magnitude (kA)

2.5

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side

800 ft

500 ft 2 300 ft 1.5 200 ft 1 150 ft 100 ft 50 ft

0.5

0 0 5 34 kV 10 15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft) 25 345 kV 500 kV

Figure 11-34 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m

11-35

Graphic Evaluation of Proposed Changes Based on Simulations

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wire 500 ft Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

3 300 ft 2 200 ft 150 ft 1 100 ft 50 ft 0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-35 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m

6 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wire

5 500 ft 4

300 ft 200 ft 150 ft 100 ft 50 ft

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-36 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m

11-36

Graphic Evaluation of Proposed Changes Based on Simulations

4.5 4 Average Phase Current Magnitude (kA) 3.5 3 2.5 2

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

500 ft

300 ft 1.5 1 0.5 0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 200 ft 150 ft 100 ft 50 ft

Figure 11-37 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m

5 4.5 Average Phase Current Magnitude (kA) 4 3.5

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wire

800 ft

500 ft 3 2.5 300 ft 2 1.5 1 0.5 0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 200 ft 150 ft 100 ft 50 ft

Figure 11-38 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m

11-37

Graphic Evaluation of Proposed Changes Based on Simulations

2.5

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Negative Side Alumoweld Shield Wire

800 ft

500 ft 1.5 300 ft 200 ft 1 150 ft 100 ft 50 ft 0.5

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 30

Figure 11-39 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m

3.5

3 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Positive Side Alumoweld Shield Wire

800 ft

2.5

500 ft

300 ft 200 ft 150 ft

1.5

100 ft 50 ft

0.5

0 0 5 34 kV 10 15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft) 25 345 kV 500 kV 30

Figure 11-40 Maximum Average Line Current for 25V Rail-to-Ground, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m

11-38

Power Pole-to-Track Distance (ft)

Power Pole-to-Track Distance (ft)

Graphic Evaluation of Proposed Changes Based on Simulations

1.6 1-mile Track Blocks 1.4 Average Phase Current Magnitude (kA) 1.2 1 0.8 200 ft 0.6 0.4 0.2 0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV 35 40 150 ft 100 ft 50 ft 300 ft
Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

500 ft Power Pole-to-Track Distance (ft) 35 Power Pole-to-Track Distance (ft) 40

Figure 11-41 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

1.6 1-mile Track Blocks 1.4 Signal Equip. Impedance = j10 ohms Average Phase Current Magnitude (kA)
Positive Side Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft

500 ft

1.2 300 ft 1 0.8 0.6 0.4 0.2 0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV

200 ft

150 ft 100 ft 50 ft

Figure 11-42 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

11-39

Graphic Evaluation of Proposed Changes Based on Simulations

1.4

1.2 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft

500 ft

0.8 300 ft 0.6 200 ft 0.4 150 ft 0.2 100 ft 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV 35

Power Pole-to-Track Distance (ft) 40 35 Power Pole-to-Track Distance (ft) 40

Figure 11-43 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

1.4

1.2 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

500 ft

0.8 300 ft 0.6 200 ft 0.4 150 ft 100 ft 0.2 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV

Figure 11-44 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

11-40

Graphic Evaluation of Proposed Changes Based on Simulations


1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft

0.8

500 ft

0.6

300 ft

0.4

200 ft 150 ft

0.2

100 ft 50 ft

0 0 5 10 138 kV 15 230 kV 20 25 345 kV 34 kV 115 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-45 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

500 ft 0.8 300 ft 0.6 200 ft 0.4 150 ft 100 ft 0.2 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-46 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

Power Pole-to-Track Distance (ft)

Power Pole-to-Track Distance (ft)

11-41

Graphic Evaluation of Proposed Changes Based on Simulations

2.5 1-mile Track Blocks


Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

Average Phase Current Magnitude (kA)

500 ft Power Pole-to-Track Distance (ft) 35 40

1.5

300 ft

0.5

200 ft 150 ft 100 ft 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV

Figure 11-47 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

2.5

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

500 ft Power Pole-to-Track Distance (ft) 35 40

1.5

300 ft 1

200 ft 0.5 150 ft 100 ft 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV

Figure 11-48 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

11-42

Graphic Evaluation of Proposed Changes Based on Simulations

1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft

0.8 500 ft 0.6 300 ft 0.4 200 ft 150 ft 0.2 100 ft 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV 35

Power Pole-to-Track Distance (ft) 40 35 Power Pole-to-Track Distance (ft) 40

Figure 11-49 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

0.8 500 ft 0.6 300 ft 0.4 200 ft 150 ft 100 ft 0.2 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV

Figure 11-50 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

11-43

Graphic Evaluation of Proposed Changes Based on Simulations


0.8 0.7 Average Phase Current Magnitude (kA) 0.6 0.5 300 ft 0.4 200 ft 0.3 0.2 0.1 0 0 5 10 138 kV 15 230 kV 20 25 345 kV 34 kV 115 kV Phase Separation (ft) 30 500 kV 35 40 150 ft 100 ft 50 ft

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) 35 Power Pole-to-Track Distance (ft) 40

500 ft

Figure 11-51 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

0.8 0.7 Average Phase Current Magnitude (kA) 0.6 0.5 0.4 0.3 0.2 0.1 0 0

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

500 ft 300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10 138 kV 115 kV

15

20 25 230 kV 345 kV Phase Separation (ft)

30 500 kV

Figure 11-52 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

11-44

Graphic Evaluation of Proposed Changes Based on Simulations

1 0.9 Average Phase Current Magnitude (kA) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

30 500 kV

35

Power Pole-to-Track Distance (ft) 40 35 Power Pole-to-Track Distance (ft) 40

Figure 11-53 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

1 0.9 Average Phase Current Magnitude (kA) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

30 500 kV

Figure 11-54 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

11-45

Graphic Evaluation of Proposed Changes Based on Simulations


0.9 0.8 Average Phase Current Magnitude (kA) 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 5 10 15 34 kV 115 kV 138 kV 230 kV 20 345 kV 25 Phase Separation (ft) 30 500 kV 35 40 300 ft 200 ft 150 ft 100 ft 50 ft 500 ft

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 35 500 kV Power Pole-to-Track Distance (ft) 40

Figure 11-55 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

0.9 0.8 Average Phase Current Magnitude (kA) 0.7 0.6 0.5 0.4

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

500 ft

300 ft 0.3 200 ft 0.2 0.1 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 150 ft 100 ft 50 ft

Figure 11-56 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

11-46

Graphic Evaluation of Proposed Changes Based on Simulations

0.8 0.7 Average Phase Current Magnitude (kA) 0.6 0.5 0.4

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40


35 500 kV Power Pole-to-Track Distance (ft) 40

500 ft

300 ft 0.3 0.2 0.1 0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 200 ft 150 ft 100 ft 50 ft

Figure 11-57 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

0.8 0.7 Average Phase Current Magnitude (kA) 0.6 0.5 0.4

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

500 ft

300 ft 0.3 0.2 0.1 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 200 ft 150 ft 100 ft 50 ft

Figure 11-58 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

11-47

Graphic Evaluation of Proposed Changes Based on Simulations


1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

0.8

500 ft

0.6

0.4

300 ft 200 ft

0.2

150 ft 100 ft 50 ft

0 0 5 10 15 25 230 kV 20 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-59 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

1 0.9 Average Phase Current Magnitude (kA) 0.8 0.7

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

500 ft 0.6 0.5 0.4 0.3 0.2 0.1 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 35 300 ft 200 ft 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 40

Figure 11-60 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

11-48

Power Pole-to-Track Distance (ft)

Graphic Evaluation of Proposed Changes Based on Simulations

1.1 1 Average Phase Current Magnitude (kA) 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

10 15 34 kV 115 kV 138 kV

230 kV 20 345 kV 25 Phase Separation (ft)

30 500 kV

35

Power Pole-to-Track Distance (ft) 40


35 Power Pole-to-Track Distance (ft) 40

800 ft

Figure 11-61 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

1 0.9 Average Phase Current Magnitude (kA) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft)

30 500 kV

Figure 11-62 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

11-49

Graphic Evaluation of Proposed Changes Based on Simulations

0.9 0.8 Average Phase Current Magnitude (kA) 0.7 0.6 0.5

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft

500 ft

300 ft 0.4 0.3 0.2 0.1 0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 35 200 ft 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 40 35 Power Pole-to-Track Distance (ft) 40

Figure 11-63 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

0.8 0.7 Average Phase Current Magnitude (kA) 0.6 0.5

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

500 ft 0.4 300 ft 0.3 200 ft 0.2 0.1 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 150 ft 100 ft 50 ft

Figure 11-64 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

11-50

Graphic Evaluation of Proposed Changes Based on Simulations


1.4

1.2 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

0.8

0.6

300 ft

0.4

200 ft 150 ft 100 ft

0.2

50 ft

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 30

Figure 11-65 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

0.8

500 ft

0.6 300 ft 0.4 200 ft 150 ft 0.2 100 ft 50 ft

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 30

Figure 11-66 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

Power Pole-to-Track Distance (ft)

Power Pole-to-Track Distance (ft)

500 ft

11-51

Graphic Evaluation of Proposed Changes Based on Simulations

1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

0.8 500 ft 0.6 300 ft 0.4 200 ft 150 ft 100 ft 50 ft

0.2

0 0 5 34 kV 10 15 230 kV 20 115 kV 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-67 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

1 0.9 Average Phase Current Magnitude (kA) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft)

345 kV

25 500 kV

Figure 11-68 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

11-52

Graphic Evaluation of Proposed Changes Based on Simulations


0.9 0.8 Average Phase Current Magnitude (kA) 0.7 0.6 0.5 300 ft 0.4 0.3 0.2 0.1 0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 30 200 ft 150 ft 100 ft 50 ft 500 ft

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft

Figure 11-69 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

0.9 0.8 Average Phase Current Magnitude (kA) 0.7 0.6 0.5

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft Power Pole-to-Track Distance (ft) 30

500 ft

300 ft 0.4 200 ft 0.3 0.2 0.1 0 0 5 34 kV 10 15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft) 25 345 kV 500 kV 150 ft 100 ft 50 ft

Figure 11-70 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

Power Pole-to-Track Distance (ft)

11-53

Graphic Evaluation of Proposed Changes Based on Simulations


1.4

1.2 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

500 ft

0.8

0.6

300 ft 200 ft 150 ft 100 ft 50 ft

0.4

0.2

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 30

Figure 11-71 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

1.4

1.2 Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

500 ft

0.8

0.6

300 ft 200 ft 150 ft 100 ft 50 ft

0.4

0.2

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV 30

Figure 11-72 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 1 Mile Blocks

11-54

Power Pole-to-Track Distance (ft)

Power Pole-to-Track Distance (ft)

Graphic Evaluation of Proposed Changes Based on Simulations

1 0.9 Average Phase Current Magnitude (kA) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft)

345 kV

25 500 kV

Figure 11-73 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

1.2

Average Phase Current Magnitude (kA)

1-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

0.8 500 ft 0.6 300 ft 0.4 200 ft 150 ft 100 ft 50 ft

0.2

0 0 5 34 kV 10 115 kV 15 230 kV 20 138 kV Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-74 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 1 Mile Blocks

11-55

Graphic Evaluation of Proposed Changes Based on Simulations

0.8 0.7 Average Phase Current Magnitude (kA) 0.6 0.5 0.4 0.3 0.2 0.1 0 0

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10 115 kV

15 230 kV 20 138 kV Vertical Phase Separation (ft)

345 kV

25 500 kV

Figure 11-75 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

1 0.9 Average Phase Current Magnitude (kA) 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

1-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10

15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft)

25 345 kV 500 kV

Figure 11-76 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 1 Mile Blocks

11-56

Graphic Evaluation of Proposed Changes Based on Simulations


0.6

Average Phase Current Magnitude (kA)

0.5

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

500 ft Power Pole-to-Track Distance (ft) 35 40 35 Power Pole-to-Track Distance (ft) 40

0.4

300 ft

0.3 200 ft 0.2 150 ft 0.1 100 ft 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 20 25 230 kV 345 kV Phase Separation (ft) 30 500 kV

Figure 11-77 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

0.6

Average Phase Current Magnitude (kA)

0.5

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

500 ft

0.4

300 ft

0.3

200 ft

0.2

150 ft

100 ft 0.1 50 ft 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

Figure 11-78 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

11-57

Graphic Evaluation of Proposed Changes Based on Simulations


0.6

Average Phase Current Magnitude (kA)

0.5

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 35 40


35 Power Pole-to-Track Distance (ft) 40

0.4

500 ft

0.3

300 ft

0.2

200 ft 150 ft

0.1

100 ft 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

Figure 11-79 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

0.6

Average Phase Current Magnitude (kA)

0.5

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

0.4

500 ft

0.3

300 ft

0.2

200 ft 150 ft

0.1

100 ft 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

Figure 11-80 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

11-58

Graphic Evaluation of Proposed Changes Based on Simulations

0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 0.25 0.2

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 35 40 35 Power Pole-to-Track Distance (ft) 40

500 ft

300 ft

200 ft 0.15 150 ft 0.1 0.05 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 100 ft 50 ft

Figure 11-81 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

0.45 0.4 Average Phase Current Magnitude (kA) 0.35

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

500 ft 0.3 0.25 0.2 0.15 0.1 0.05 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 300 ft

200 ft 150 ft 100 ft 50 ft

Figure 11-82 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

11-59

Graphic Evaluation of Proposed Changes Based on Simulations


0.9 0.8 Average Phase Current Magnitude (kA) 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 35 40 200 ft 150 ft 100 ft 50 ft 300 ft

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

500 ft Power Pole-to-Track Distance (ft) 35 Power Pole-to-Track Distance (ft) 40

Figure 11-83 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

0.9 0.8 Average Phase Current Magnitude (kA) 0.7 0.6 0.5

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

500 ft

300 ft 0.4 0.3 0.2 0.1 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

200 ft 150 ft 100 ft 50 ft

Figure 11-84 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

11-60

Graphic Evaluation of Proposed Changes Based on Simulations


0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 35 40 300 ft

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft

500 ft

200 ft 150 ft 100 ft 50 ft

Figure 11-85 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 0.25

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

500 ft

300 ft 0.2 0.15 0.1 0.05 0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV 35 200 ft 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 40

Figure 11-86 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

Power Pole-to-Track Distance (ft)

11-61

Graphic Evaluation of Proposed Changes Based on Simulations

0.3

Average Phase Current Magnitude (kA)

0.25

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) 35 40 35 Power Pole-to-Track Distance (ft) 40

0.2

500 ft

300 ft 0.15 200 ft 0.1 150 ft 100 ft 0.05 50 ft

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

Figure 11-87 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

0.3

Average Phase Current Magnitude (kA)

0.25

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

500 ft 0.2 300 ft 0.15 200 ft 150 ft 0.1 100 ft 50 ft 0.05

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

Figure 11-88 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Vertical Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

11-62

Graphic Evaluation of Proposed Changes Based on Simulations


0.4 0.35 Average Phase Current Magnitude (kA) 0.3 0.25 0.2 0.15 0.1 0.05 0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 35 40 300 ft 500 ft

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

200 ft 150 ft 100 ft 50 ft

Figure 11-89 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

0.4 0.35 Average Phase Current Magnitude (kA) 0.3 0.25 0.2 0.15 0.1 0.05 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft)

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

30 500 kV

35

Power Pole-to-Track Distance (ft) 40

Figure 11-90 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

Power Pole-to-Track Distance (ft)

11-63

Graphic Evaluation of Proposed Changes Based on Simulations

0.35

0.3 Average Phase Current Magnitude (kA)

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40 35 500 kV Power Pole-to-Track Distance (ft) 40

0.25

0.2

500 ft

0.15 300 ft 0.1 200 ft 150 ft 100 ft 50 ft

0.05

0 0 5 34 kV 10 15 115 kV138 kV 230 kV 20 345 kV 25 Phase Separation (ft) 30

Figure 11-91 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

0.35

0.3 Average Phase Current Magnitude (kA)

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

0.25

0.2

500 ft

0.15 300 ft 0.1 200 ft 150 ft 100 ft 50 ft

0.05

0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

Figure 11-92 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

11-64

Graphic Evaluation of Proposed Changes Based on Simulations


0.3

Average Phase Current Magnitude (kA)

0.25

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40 35 500 kV Power Pole-to-Track Distance (ft) 40

0.2 500 ft 0.15 300 ft 0.1 200 ft 150 ft 100 ft 50 ft 0.05

0 0 5 34 kV 10 15 230 kV 20 25 345 kV 115 kV138 kV Phase Separation (ft) 30

Figure 11-93 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

0.3

Average Phase Current Magnitude (kA)

0.25

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

0.2 500 ft 0.15 300 ft 0.1 200 ft 150 ft 100 ft 50 ft 0.05

0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

Figure 11-94 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

11-65

Graphic Evaluation of Proposed Changes Based on Simulations

0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 0.25 0.2

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

500 ft

300 ft 0.15 200 ft 0.1 0.05 0 0 5 10 15 230 kV 20 25 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 35 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 40 35 Power Pole-to-Track Distance (ft) 40

Figure 11-95 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

0.4 0.35 Average Phase Current Magnitude (kA) 0.3 0.25 0.2 0.15 0.1 0.05 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft)

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

30 500 kV

Figure 11-96 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

11-66

Graphic Evaluation of Proposed Changes Based on Simulations


0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 0.25 500 ft 0.2 0.15 0.1 0.05 0 0 5 34 kV 10 15 115 kV138 kV 230 kV 20 345 kV 25 Phase Separation (ft) 30 500 kV 35 40 300 ft 200 ft 150 ft 100 ft 50 ft

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

Figure 11-97 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

0.4 0.35 Average Phase Current Magnitude (kA) 0.3

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft 0.25 0.2 0.15 300 ft 0.1 0.05 0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30 500 kV 35 200 ft 150 ft 100 ft 50 ft

500 ft

Power Pole-to-Track Distance (ft) 40

Figure 11-98 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

Power Pole-to-Track Distance (ft)

800 ft

11-67

Graphic Evaluation of Proposed Changes Based on Simulations


0.35

0.3 Average Phase Current Magnitude (kA)

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft

0.25 500 ft 0.2 300 ft 200 ft 0.1 150 ft 100 ft 50 ft 0.05

0.15

0 0 5 34 kV 10 15 230 kV 20 25 345 kV 115 kV138 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-99 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

0.3

Average Phase Current Magnitude (kA)

0.25

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) 35 500 kV 40

0.2 500 ft 0.15 300 ft 0.1 200 ft 150 ft 100 ft 50 ft

0.05

0 0 5 10 15 20 25 230 kV 345 kV 34 kV 115 kV 138 kV Phase Separation (ft) 30

Figure 11-100 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

11-68

Power Pole-to-Track Distance (ft)

Graphic Evaluation of Proposed Changes Based on Simulations


0.6

Average Phase Current Magnitude (kA)

0.5

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

0.4

500 ft

0.3 300 ft 0.2 200 ft 150 ft 0.1 100 ft 50 ft 0 0 5 34 kV 10 115 kV 138 kV 15 230 kV 20 Vertical Phase Separation (ft) 345 kV 25 500 kV

Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

Figure 11-101 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10 115 kV 138 kV 15 230 kV 20 Vertical Phase Separation (ft)

345 kV

25 500 kV

Figure 11-102 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

11-69

Graphic Evaluation of Proposed Changes Based on Simulations


0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 500 ft 0.25 0.2 0.15 0.1 0.05 0 0 5 34 kV 10 115 kV 138 kV15 230 kV 20 Vertical Phase Separation (ft) 345 kV 25 500 kV 30

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft)

300 ft 200 ft 150 ft 100 ft 50 ft

Figure 11-103 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

0.4 0.35 Average Phase Current Magnitude (kA) 0.3 0.25 0.2

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

500 ft

300 ft 0.15 0.1 0.05 0 0 5 34 kV 10 15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft) 25 345 kV 500 kV 200 ft 150 ft 100 ft 50 ft

Power Pole-to-Track Distance (ft) 30

Figure 11-104 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

11-70

Graphic Evaluation of Proposed Changes Based on Simulations

0.35

0.3 Average Phase Current Magnitude (kA)

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side

800 ft Power Pole-to-Track Distance (ft) 30


Power Pole-to-Track Distance (ft) 30

0.25 500 ft 0.2 300 ft 0.15 200 ft 150 ft 0.1 100 ft 50 ft 0.05

0 0 5 34 kV 10 15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft) 25 345 kV 500 kV

Figure 11-105 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

0.35

0.3 Average Phase Current Magnitude (kA)

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side

800 ft

0.25 500 ft 0.2 300 ft 0.15 200 ft 0.1 150 ft 100 ft 50 ft

0.05

0 0 5 34 kV 10 115 kV 138 kV 15 230 kV 20 Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-106 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, Without Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

11-71

Graphic Evaluation of Proposed Changes Based on Simulations


0.5 0.45 Average Phase Current Magnitude (kA) 0.4 0.35 0.3 0.25 300 ft 0.2 0.15 0.1 0.05 0 0 5 34 kV 10 115 kV 138 kV 15 230 kV 20 Vertical Phase Separation (ft) 345 kV 25 500 kV 30 200 ft 150 ft 100 ft 50 ft

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

Figure 11-107 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

0.6

Average Phase Current Magnitude (kA)

0.5

2-mile Track Blocks Earth Resistivity = 10 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

0.4

500 ft

0.3 300 ft 0.2 200 ft 0.1 150 ft 100 ft 50 ft

0 0 5 34 kV 10 115 kV 138 kV 15 230 kV 20 Vertical Phase Separation (ft) 345 kV 25 500 kV 30

Figure 11-108 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 10 ohm-m, 2 Mile Blocks

11-72

Power Pole-to-Track Distance (ft)

Power Pole-to-Track Distance (ft)

500 ft

Graphic Evaluation of Proposed Changes Based on Simulations


0.4 0.35 Average Phase Current Magnitude (kA) 0.3 0.25 500 ft 0.2 0.15 0.1 0.05 0 0 5 34 kV 10 115 kV 138 kV15 230 kV 20 Vertical Phase Separation (ft) 345 kV 25 500 kV 30 300 ft 200 ft 150 ft 100 ft 50 ft

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) Power Pole-to-Track Distance (ft) 30

Figure 11-109 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

0.45 0.4 Average Phase Current Magnitude (kA) 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 0

2-mile Track Blocks Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

500 ft

300 ft 200 ft 150 ft 100 ft 50 ft

5 34 kV

10

15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft)

25 345 kV 500 kV

Figure 11-110 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 100 ohm-m, 2 Mile Blocks

11-73

Graphic Evaluation of Proposed Changes Based on Simulations


0.3

Average Phase Current Magnitude (kA)

0.25

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Negative Side Alumoweld Shield Wire

800 ft Power Pole-to-Track Distance (ft) 30 Power Pole-to-Track Distance (ft) 30

0.2

500 ft

0.15

300 ft 200 ft

0.1

150 ft 100 ft 50 ft

0.05

0 0 5 34 kV 10 15 20 115 kV 138 kV 230 kV Vertical Phase Separation (ft) 25 345 kV 500 kV

Figure 11-111 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Negative Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

0.35

0.3 Average Phase Current Magnitude (kA)

2-mile Track Blocks Earth Resistivity = 1000 ohm-m Ballast Resistivity = 100 ohm-kft Signal Equip. Impedance = j10 ohms Positive Side Alumoweld Shield Wire

800 ft

0.25

500 ft

0.2

300 ft 200 ft 150 ft

0.15

0.1

100 ft 50 ft

0.05

0 0 5 34 kV 10 115 kV 138 kV 15 230 kV 20 Vertical Phase Separation (ft) 345 kV 25 500 kV

Figure 11-112 Maximum Average Line Current for 5V Rail-to-Rail with Shorted IJ, Horizontal Delta Circuit with Unbalance, Positive Side, With Shield Wire, Earth Resistivity 1000 ohm-m, 2 Mile Blocks

11-74

Graphic Evaluation of Proposed Changes Based on Simulations


60

Current Through Person from Signal Pole Line Conductor (mA/mi)

Negative Side 50 ft

50 Power Pole-to-Track Distance (ft) 35

40
Electrostatically Induced Current

30

65 ft

20

10

80 ft 150 ft 125 ft

0 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft)

100 ft 30 500 kV 40

Figure 11-113 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Negative Side, Without Shield Wire

Current Through Person from Signal Pole Line Conductor (mA/mi)

14 Positive Side 12 50 ft Power Pole-to-Track Distance (ft) 35

10
Electrostatically Induced Current

8
I

125 ft 150 ft 100 ft

4 65 ft 80 ft

0 0 5 34 kV 10 138 kV 15 230 kV 20 25 345 kV 115 kV Phase Separation (ft) 30 500 kV 40

Figure 11-114 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Positive Side, Without Shield Wire

11-75

Graphic Evaluation of Proposed Changes Based on Simulations

Current Through Person from Signal Pole Line Conductor (mA/mi)

60

Negative Side Alumoweld Shield Wire

50 ft

50 Power Pole-to-Track Distance (ft) 35

40
Electrostatically Induced Current

30

65 ft

20 80 ft 10 150 ft 125 ft 100 ft 0 5 34 kV 10 138 kV 115 kV 15 230 kV 20 25 345 kV Phase Separation (ft) 30 500 kV

40

Figure 11-115 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Negative Side, With Shield Wire

Current Through Person from Signal Pole Line Conductor (mA/mi)

16 Positive Side Alumoweld Shield Wire 14 12 10


Electrostatically Induced Current

50 ft

8 6 4 2

65 ft 150 ft 125 ft 100 ft 80 ft

0 0 5 34 kV 10 138 kV 15 230 kV 20 25 345 kV 115 kV Phase Separation (ft) 30 500 kV 35 40

Figure 11-116 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Vertical Circuit, Positive Side, With Shield Wire

11-76

Power Pole-to-Track Distance (ft)

Graphic Evaluation of Proposed Changes Based on Simulations

Current Through Person from Signal Pole Line Conductor (mA/mi)

120

Negative Side

100

50 ft Power Pole-to-Track Distance (ft) 35 500 kV

80
Electrostatically Induced Current

65 ft

60

80 ft

40

100 ft 125 ft 150 ft

20

0 0 5 34 kV 10 15 230 kV 20 345 kV 25 115 kV 138 kV Phase Separation (ft) 30 40

Figure 11-117 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Negative Side, Without Shield Wire
120

Current Through Person from Signal Pole Line Conductor (mA/mi)

Positive Side

100

50 ft Power Pole-to-Track Distance (ft) 35 500 kV

80
Electrostatically Induced Current

65 ft

60

80 ft

40

100 ft 125 ft 150 ft

20

0 0 5 34 kV 10 15 25 230 kV 20 345 kV 115 kV 138 kV Phase Separation (ft) 30 40

Figure 11-118 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Positive Side, Without Shield Wire

11-77

Graphic Evaluation of Proposed Changes Based on Simulations


120

Current Through Person from Signal Pole Line Conductor (mA/mi)

Negative Side Two Alumoweld Shield Wires 50 ft Power Pole-to-Track Distance (ft) 35 500 kV

100

80
Electrostatically Induced Current

65 ft

60

80 ft

40

100 ft 125 ft 150 ft

20

0 0 5 34 kV 10 15 20 25 230 kV 345 kV 115 kV 138 kVPhase Separation (ft) 30 40

Figure 11-119 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Negative Side, With Shield Wire

Current Through Person from Signal Pole Line Conductor (mA/mi)

120

Positive Side Two Alumoweld Shield Wires 50 ft Power Pole-to-Track Distance (ft) 35 500 kV

100

80
Electrostatically Induced Current

65 ft

60

80 ft

40

100 ft 125 ft 150 ft

20

0 0 5 34 kV 10 15 20 25 230 kV 345 kV 115 kV 138 kV Phase Separation (ft) 30 40

Figure 11-120 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Circuit, Positive Side, With Shield Wire

11-78

Graphic Evaluation of Proposed Changes Based on Simulations


70

Current Through Person from Signal Pole Line Conductor (mA/mi)

Negative Side 50 ft Power Pole-to-Track Distance (ft) 30

60

50

40
I

Electrostatically Induced Current

65 ft

30

80 ft 100 ft 125 ft

20

10

150 ft

0 0 5 34 kV 10 15 20 138 kV 230 kV 345 kV 115 kV Vertical Phase Separation (ft) 25 500 kV 35

Figure 11-121 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Negative Side, Without Shield Wire

Current Through Person from Signal Pole Line Conductor (mA/mi)

70

Positive Side

60

50 ft Power Pole-to-Track Distance (ft) 30

50 65 ft 40
I
Electrostatically Induced Current

80 ft 30

100 ft 20 125 ft 10 150 ft

0 0 5 34 kV 10 15 20 138 kV 230 kV 345 kV 115 kV Vertical Phase Separation (ft) 25 500 kV 35

Figure 11-122 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Positive Side, Without Shield Wire

11-79

Graphic Evaluation of Proposed Changes Based on Simulations


70

Current Through Person from Signal Pole Line Conductor (mA/mi)

Negative Side Alumoweld Shield Wire 50 ft Power Pole-to-Track Distance (ft) 30

60

50

40

Electrostatically Induced Current

65 ft

30

80 ft

20

100 ft 125 ft 150 ft

10

0 0 5 34 kV 10 15 20 138 kV 230 kV 345 kV 115 kV Vertical Phase Separation (ft) 25 500 kV 35

Figure 11-123 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Negative Side, With Shield Wire

Current Through Person from Signal Pole Line Conductor (mA/mi)

70 Positive Side Alumoweld Shield Wire 60 50 ft Power Pole-to-Track Distance (ft) 30

50 65 ft 40
Electrostatically Induced Current

80 ft

30

100 ft 20 125 ft 10 150 ft

0 0 5 34 kV 10 15 20 138 kV 230 kV 345 kV 115 kV Vertical Phase Separation (ft) 25 500 kV 35

Figure 11-124 Current to Person Touching Signal Pole Line Capacitively Coupled to Power Line for a Horizontal Delta Circuit, Positive Side, With Shield Wire

11-80

Graphic Evaluation of Proposed Changes Based on Simulations

10000 10-ohm earth resistivity 155-lb person EHSS OHSW at 87.7ft high Lowest phase conductor at 40ft high Allowable Fault Current (Amps)

3 cycle 6 cycle 1000 10 cycle 15 cycle 20 cycle

100 10 100 Power-to-Track Distance (ft) 1000

Figure 11-125 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 155-lb Person, Earth Resistivity 10 ohm-m

10000 10-ohm earth resistivity 110-lb person EHSS OHSW at 87.7ft high Lowest phase conductor at 40ft high Allowable Fault Current (Amps)

3 cycle 1000 6 cycle 10 cycle 15 cycle 20 cycle

100 10 100 Power-to-Track Distance (ft) 1000

Figure 11-126 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 110-lb Person, Earth Resistivity 10 ohm-m

11-81

Graphic Evaluation of Proposed Changes Based on Simulations

10000

100-ohm earth resistivity 155-lb person EHSS OHSW at 87.7ft high Lowest phase conductor at 40ft high

Allowable Fault Current (Amps)

3 cycle 1000 6 cycle 10 cycle 15 cycle 20 cycle

100 10 100 Power-to-Track Distance (ft) 1000

Figure 11-127 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 155-lb Person, Earth Resistivity 100 ohm-m

10000 100-ohm earth resistivity 110-lb person EHSS OHSW at 87.7ft high Lowest phase conductor at 40ft high Allowable Fault Current (Amps)

1000

3 cycle 6 cycle 10 cycle 15 cycle 20 cycle

100 10 100 Power-to-Track Distance (ft) 1000

Figure 11-128 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 110-lb Person, Earth Resistivity 100 ohm-m

11-82

Graphic Evaluation of Proposed Changes Based on Simulations

10000 1000-ohm earth resistivity 155-lb person EHSS OHSW at 87.7ft high Lowest phase conductor at 40ft high Allowable Fault Current (Amps)

3 cycle 6 cycle 10 cycle 15 cycle 20 cycle

1000

100 10 100 Power-to-Track Distance (ft) 1000

Figure 11-129 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 155-lb Person, Earth Resistivity 1000 ohm-m

10000 1000-ohm earth resistivity 110-lb person EHSS OHSW at 87.7ft high Lowest phase conductor at 40ft high Allowable Fault Current (Amps)

3 cycle 6 cycle 1000 10 cycle 15 cycle 20 cycle

100 10 100 Power-to-Track Distance (ft) 1000

Figure 11-130 Allowable Fault Current Vs. Power-to-Track Distance for Safe Track Touch Potential, 110-lb Person, Earth Resistivity 1000 ohm-m

11-83

Graphic Evaluation of Proposed Changes Based on Simulations

500
3/8" HSS OHSW

Rail Arrester Current (amps/kA fault current)

400

300

200
Soil Resistivity Ohm-m

1000

100
100

0 10 100 1000 10000 Distance Power to Track (feet)

Figure 11-131 Calculated Single Phase to Ground Fault-Induced Rail Current for Long Exposure

120

Rail Arrester Current (percent of long exposure faultinduced rail current)

Ballast Resistivity = 3 ohm-m Power Pole to Track Distance = 100 ft 3/8" HSS OHSW

100

80

60 10 40 Soil Resistivity Ohm-m 20 1000 100

0 0.1 1 10 100

Exposure Length (miles)

Figure 11-132 Fault-Induced Rail Current Vs. Exposure Length

11-84

12
COMPUTER MODELING OF JOINT CORRIDORS
This chapter provides in depth guidance as to when and how to apply computer modeling to evaluate potential ac interference issues. Mr. Frazier has over 30 years of experience in the management and conduct of applied research programs and the application of the research results to the solution of system-specific problems. His major fields of experience include electromagnetic effects, electromagnetic compatibility, cathodic-protection systems, and detection systems. He is the author of the EPRI computer program CORRIDOR that has been extensively used by the engineering community to predict power-line interference on co-located structures. He is nationally known for his work on the electromagnetic compatibility of electric power, pipeline, and railroad systems. As President of Corr Comp Co., Mr. Frazier provides services to the power/pipeline/railroad industries in the pursuit of compatible common utility corridors. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE.

12-1

Computer Modeling of Joint Corridors

Introduction
Computer modeling of joint railroad and power system corridors is a very powerful tool for understanding how ac interference from power lines can affect railroad systems and electrical safety. Most railroads and power companies will hire consultants to perform the computer modeling, rather than retain the expertise in house. Also, each software package for computer modeling requires specific training and experience to use reliably. For these reasons this chapter is geared toward preparing the reader to identify the inputs required to perform the modeling, and specify the outputs and recommendations the consultant should provide. If the software and expertise are available within the organization, then all of the same information still applies. However, training in the use of specific software must be obtained from the vendor. The quick-start version of this chapter presents the very basic requirements for retaining a consulting engineer and completing a useful computer modeling study. The detailed version of the chapter provides a comprehensive tutorial on how to perform a study based on computer modeling.

Quick-Start Version
What Computer Modeling Will and Will Not Do Computer modeling allows us to evaluate what will happen in a worst-case situation. Normally it is impossible to measure worst-case conditions because it is unacceptable to create those conditions on operational systems. On the other hand, most computer modeling is restricted to magnetic induction and electric induction1. Conducted interference is difficult to accurately model because of the difficulty in defining the paths. So, computer modeling is usually applied in two cases: A proposed new joint corridor, or proposed changes to an existing corridor, for the purpose of minimizing interference problems from magnetic and electric induction through careful design and appropriate mitigation An existing joint corridor for the purpose of determining if existing interference levels are consistent with magnetic and electric induction, and to determine appropriate mitigation

General The following recommendations can be followed as a general approach for each project. Specific requirements can be developed for each project as necessary.

The old term inductive coordination refers to magnetic induction and electric induction only. Conducted interference was usually not considered.

12-2

Computer Modeling of Joint Corridors

It is extremely important to develop and maintain a cooperative working relationship between the railroad and the power company. Whenever possible they need to work together to get mutually beneficial results. The consultant performing the study should complete a checklist such as the following: Conduct a project initiation meeting, including all parties, to determine the project requirements, schedule, approach, interfaces, and tasks. Obtain the required information including power system and railroad facilities and any other facilities (such as pipelines) that could affect the situation. Establish mutually agreeable, tolerable levels of ac interference in the railroad facilities. Evaluate the magnitude of induced voltages in the railroad facilities under steady state and fault conditions. Develop alternate designs and/or mitigation options that limit induction to tolerable levels. Note: These evaluations will quantify the effectiveness and relative performance of the alternatives considered. Prepare comparative cost estimates for the developed alternatives and present them in a preliminary report to the power company and the railroad. Review the report with all parties to determine the best and most economical solution. Prepare a final report detailing the results of the study and emphasizing the mutually agreedupon solution.

The consultant may be required to testify as to the reliability of the analysis before a state commission or in court. Required Information The electromagnetic compatibility information form, Figures 12-1 through 12-4, meets the informational requirements of most projects. Sections II and III are power company data. Section IV is railroad data. It is important to perform this study early in the siting/design process so that changes in siting and system design can be made at minimal cost.

12-3

Computer Modeling of Joint Corridors

Electromagnetic Compatibility Information


Transmission Lines Paralleling Railroad Communication and Signal Circuits I General Date Supplied ________

Power Company __________________________________ Project _________________________ Railroad Company _____________________________ Consultant _______________________ Location Description _____________________________________________________________ Designation of Lines _________ _________ _________ _________

Mile Posts (from/to) - End points of exposure are considered to be where the transmission line first comes within 200 feet of the nearest track. - (circle one: Proposed/Existing) ____/____P E ____/____P E ____/____P E ____/____P E

II

Power - Electrical Conductor Size Type ___________ ___________ ___________ ___________ ___________

Date Supplied ________ ___________ ___________ ___________ ___________ ___________ ___________

Diameter (specify units) ___________

Number of subconductors per phase and separation (specify units) ___________ ___________ ___________ ___________

Final Sag (specify units) at _____ degrees ___ (C or F) (typically 120 deg. F) ___________ ___________ ___________ ___________

Shield Wire Size Type

___________ ___________

___________ ___________ ___________

___________ ___________ ___________

___________ ___________ ___________

Diameter (specify units) ___________ Geometric Mean Radius (specify units) ___________

___________

___________

___________

Final Sag (specify units) at _____ degrees ___ (C or F) (typically 90 deg. F) ___________ ___________ ___________ ___________

AC impedance per mile at _____ degrees C (typically 25 deg. C) ___________ ___________ ___________ ___________

Figure 12-1 Electromagnetic Compatibility Information Form, Page 1 of 4

12-4

Computer Modeling of Joint Corridors

II.

Power-Electrical (Continued) Shield Wire (Continued) Grounding Continuous or Segmented (identify segments on plan drawing) ___________ Ruling Span (specify units) ___________ Resistance of structure grounds ___________ Substations Name Location Ground Resistance ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Nominal Voltage(phase-to-phase) Load (MVA) Initial Average Max. Steady State Ultimate Average Max. Steady State

___________

___________

___________ ___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Phase Current (Amperes) Initial Average Max. Steady State Ultimate Average Max. Steady State ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Figure 12-2 Electromagnetic Compatibility Information Form, Page 2 of 4

12-5

Computer Modeling of Joint Corridors

II.

Utility-Electrical (Continued) Maximum Load Current in Amperes during single-circuit operation of multiple circuit lines: ___________ Direction of Power Flow ___________ ___________ ___________ ___________ ___________ ___________ ___________

Fault Current - Maximum for fault to ground at ends of exposure (RMS Amps. Sym.) End ____________ 1 Phase 3 Phase End ____________ 1 Phase 3 Phase ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Clearing time with Primary Relay (sec.) ___________ Maximum # of Reclosing Attempts ___________ ___________ ___________ ___________ ___________ ___________ ___________

Clearing time with Secondary Relay (sec.) ___________ ___________ ___________ ___________

Division of fault current between overhead ground wires and earth: ___________ Estimated earth resistivity Estimated length of exposure Proposed Existing ___________ ___________ ___________ ___________ Ohm-meters ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Figure 12-3 Electromagnetic Compatibility Information Form, Page 3 of 4

12-6

Computer Modeling of Joint Corridors


III Utility - Mechanical Date Supplied ________

Section drawings showing: Dimensions and positions of all phase wires and overhead ground wires relative to the structure centerline. (For the purpose of inductive coordination studies, conductor heights are assumed to be the attachment height minus 2/3 * final sag at temperature.) The horizontal distance between the centerline of power structures, the centerline of railroad pole line, and centerline of main track throughout the exposure. Line number and phase designation for each conductor. Elevation differences between railroad and transmission facilities. Schematic drawing of the power system in the area near the exposure showing the paralleling lines, locations of all transposition points, substations and location of all neutral grounding points throughout and near the ends of the exposure. IV Railroad Date Supplied ________

Drawings showing the railroad tracks, locations of insulating joints, the resistance between the rails, whether the rails are welded or bolted, the rail weight, and the locations of any existing ground mats or ground rods installed by railroad. Drawings showing the details of the signal circuits, their associated equipment (relays, batteries, etc.), the locations of the signal and communication circuits, whether the signal circuits are installed in open wire lines or buried, the wire size, length, and the number of all signal and communication wires, and all details regarding their physical location, the structures supporting the wires, the locations of any transpositions, and the details of any existing shielding scheme. Specific data on the existing surge protection facilities including the use of carbon blocks, gas tubes, etc. and their grounding method and resistance. Drawings showing the details of communication circuits such as telephone lines and their signal levels and frequencies. Recommended induced voltage tolerable limits if different from the AAR/EEI joint agreement. Information on the susceptibility of your signal equipment to 60 Hz voltages if such information is available. (If not, we will use values published in the technical literature for similar signal equipment.) Anticipated future plans that would effect any of the above and details as to what would change.

Figure 12-4 Electromagnetic Compatibility Information Form, Page 4 of 4

Results The preliminary and final reports could include, but not be limited to the following: Executive Summary Description of railroad and power company facilities Description of steady-state, switching, and fault currents used in the analysis 12-7

Computer Modeling of Joint Corridors

Clear statement of levels of ac interference to achieve through design and mitigation Procedure used for the computer modeling analysis Details results of the analysis Summary of the results Mitigation options including effectiveness and cost in the preliminary report. The final report should emphasize the mutually agreed-upon solution. Mitigation recommendations, including installation tolerances and their effect on performance Estimated cost of recommended systems Any drawings necessary for clarification Field test reports if applicable References and support documentation

Choosing a Consultant Care should be used in choosing a consultant. Many excellent engineering firms can provide the needed services. Make sure that the consultants used are well versed in railroad systems, power systems, and electromagnetic compatibility. Also, be sure that the software used has been proven through validation with measured results. Choosing Software The software used can be widely available such as CORRIDOR and CDEGS, or it can be a proprietary software program. In any case it is important that the software used has been proven through validation with measured results. There have been cases of inaccurate results through faulty software.

Detailed Version
Overview of Power System-Railroad Interactions The purpose of modeling of railroad and power systems is most often to identify if personnel safety or signal-system compatibility problems are likely to occur if a new proposed power system becomes operational. Modeling can also be useful to assist in the design of measures to help reduce the severity of coupled energy from an existing or proposed power system to a railroad system. The goal of mitigation design is to achieve compatible operation of the two systems. Modeling can also be useful in helping to understand an existing interference situation, so that steps can be taken to reduce the adverse influence of the power system on the railroad signal system.

12-8

Computer Modeling of Joint Corridors

The goal of this chapter is not to delve into the theoretical aspects of the problem that are necessary to develop a computer program suitable for analyzing the interaction between power systems and railroads. Rather, a goal of this chapter is to pass on to the reader an awareness of how some of the important parameters of both systems influence the voltage and current that is coupled into the railroad system. Another goal of this chapter is to illustrate procedures that have been used to conduct studies of planned common corridors, to serve as a guideline for structuring future studies. For most cases of concern the interaction of a power line and a rail-system is caused by the power line running nominally parallel to the rail line for some distance. How far the power line and the rail line need to be nominally parallel to result in possible compatibility problems can not be simply answered. There are many factors that contribute to the answer. One of the objectives of this chapter is to illustrate some of the more important factors that contribute to determining how the two systems interact. That is, this chapter will describe some important properties of the power system and the rail system that influence the voltages and currents that are coupled into the rail system, and whether or not these induced levels should be a concern. The preceding chapter, Chapter 11, presents some simple graphical approaches that can be uses to provide a preliminary estimate of the induced rail-system voltage. Power Systems Steady State Magnetic Field for Railroad Compatibility Studies Quantifying Magnetic Field Coupling A power line that is roughly parallel to a rail line causes a distributed voltage to be induced into the rails or other rail-system conductors by the current that flows in the power line conductors. The term distributed voltage implies that the longer the length of rail that is excited, the larger the induced voltage. The voltage per unit length of exposure (the induced voltage per meter, or per foot, or per 1000 feet) is related to the current in the power line conductors by an expression that depends on the geometry (the separation between the power conductors and the rail conductor) the electrical properties of the earth, and the frequency of the power line current. The principal frequency of the power line current is 60 Hz (or 50 Hz), although the current may also have some harmonic components, which are integer multiples of 60 Hz (or 50 Hz). The more formal term for the induced voltage per unit length is the longitudinal electric field (LEF). The expression that relates the longitudinal electric field and the power line current is called transfer impedance.

12-9

Computer Modeling of Joint Corridors

What is the Longitudinal Electric Field (LEF)? When modeling magnetically induced voltages on railroad tracks (inductive coordination), we usually calculate the transfer impedance, Z. When we multiply the current in the source conductor by this transfer impedance we get the longitudinal electric field (LEF) imposed on the rail. LEF is represented by the symbol E and has units of volts per unit length. This is a confusing concept. The main source of confusion comes directly from the names we give these quantities. We all know that the coupling paths are magnetic induction, electric induction, and conduction. The computer modeling we are addressing in this chapter addresses only magnetic induction. We know that the physical process by which magnetic induction happens is that the current in the source conductor creates a magnetic field; that this magnetic field falls across the rail; that this magnetic field pushes on electrons in the rail; that this pushing, or force, causes the voltage (a.k.a. electromotive force) we see induced on the rail. So, where does the electric field in longitudinal electric field enter into the process? In fact, the electric field, as we have been using the term in Chapter 2, is not what we are talking about. We are encountering confusion brought on by the use of the same terminology in two different areas of specialization. The electric field we use in this chapter exists only in the rail. It is the voltage gradient in the rail measured per unit length. We normally think of electric fields as something that exists in the space between conductors. Since both uses of the term are valid, this tends to cause confusion. For that reason we will always try to use the full name of longitudinal electric field (LEF) to describe fields within the rail. A similar problem exists when we show the direction of current flow in a simple circuit. A very long time ago, before we knew electrons existed, the flow of current was observed in a conductor between terminals of differing electric charge. These terminals were randomly described as positive and negative. And, the direction of current flow was defined as from positive to negative. It was many years later that we discovered that it is the negatively charged electrons that actually move. That leaves us in the conflicted condition of describing the movement of particles in one direction as current in the opposite direction. Our problem with longitudinal electric field is very similar. Voltage, such as that induced on the rail, is sometimes referred to as electromotive force (units of volts). In describing a field effect (in volts per unit length) that causes the voltage, we use longitudinal electromotive force or longitudinal electric field. It is really that simple. We would be far better off if our predecessors had chosen a different name for this quantity, but they didnt. In short, these are two different ways of describing the same phenomena. One is literal: current magnetic field force on electrons voltage The other is analogous: current times the transfer impedance longitudinal electric field voltage The physical mechanism is the same (magnetic field induction). Both methods give the same result. When modeling, particularly with computers, there are real advantages to using the transfer impedance method. The practical challenge when implementing the transfer impedance method is to develop the impedance matrix accurately. The use of longitudinal electromotive force or longitudinal electric field in setting standards for EMI is very helpful. The ability of the magnetic field from a power line to induce voltage in a receptor can be measured using the longitudinal electric field in volts per unit length. This is a value to which the power company can design. It is independent of the design of the railroad facilities.

12-10

Computer Modeling of Joint Corridors

Sometimes the voltage that is induced into a rail is discussed in terms of the magnetic field, which is caused by the current in the power-line conductors (as in Chapter 2). Use of the magnetic field is helpful in visualizing the coupling between current-carrying power-line wires and nearby conductors such as railroad tracks. The magnetic field and the longitudinal electric field are related by a mathematical transformation. For the purpose of analyzing the coupling between power line current-carrying conductors and nearby rails, the transfer impedance method is almost always used (see sidebar: What is the Longitudinal Electric Field?). The transfer impedance method of analysis relates the current I in a power conductor the longitudinal electric field E at an observation point by a simple ohms law-type relationship: E = Zt I
Eq. 12-1

Figure 12-5 shows the geometry for a current carrying conductor and an observation point, where h is the height of the current caring conductor, and h is the height of the observation point where E is to be determined. If the observation point is a rail conductor, which is at the surface of the earth, h = 0. The original derivation of the transfer impedance in a form that could be readily evaluated was by Carson in 1926 [76]. Carsons expression for the transfer impedance (also termed the mutual impedance) Zt can be evaluated by summing an infinite series. One of the difficulties with the infinite series form derived by Carson was that it is difficult to visualize the importance of the various terms in the rather complex expression.

Figure 12-5 Geometry for Conductor and Observation Point Above the Earth

12-11

Computer Modeling of Joint Corridors

A more easily visualized expression for the mutual impedance has been derived, which considers an equivalent return conductor at a complex distance below the surface of the earth [77] and is summarized by Olsen [78]. Figure 12-6 is a sketch that shows a current carrying conductor, the observation point and an image-current-carrying conductor below the earths surface. An approximate expression for the transfer impedance is,

Z c = K ln
where,

D" D'

Eq. 12-2

K = jf

j = 1
f is the frequency in Hertz

= 4 10 7 henry/meter, the magnetic permeability of free space


D ' = (h h') + d 2 2 is the distance from the current carrying conductor to the observation point (in meters), as is illustrated in Figure 12-6.
2

D" = (h + h'+2 p ) + d 2 2 is the distance from the image of the current carrying conductor to the observation point (in meters), as is illustrated in Figure 12-6.
2

p=

j 2 f

is the complex depth of the image of the earth surface.

is the earth resistivity (ohmm).

12-12

Computer Modeling of Joint Corridors

Figure 12-6 Complex Image Plane Geometry

The current in the image conductor flows in the opposite direction from the current in the actual power-line conductor. Therefore the contribution from the image conductor tends to cancel the field contribution from the actual conductor. For an ideal zero-resistance earth, D becomes approximately the same as D and the transfer impedance in Eq. 12.2 goes to zero. However for non-zero resistivity earth, D becomes larger than D. Higher values of earth resistivity result in deeper image conductor depths and larger values for the image-conductor to observation point distance D. As the distance to the image conductor becomes larger less cancellation occurs, and the magnitude of the transfer impedance Zt becomes larger. Thus, higher values of earth resistivity results in higher values of the transfer impedance. Which means that for a given geometry and power line current, the induced rail voltage can be higher for higher values of earth resistivity. Figure 12-7 shows the approximate relationship between the depth of the image conductor and the soil resistivity, which indicates that the soil resistivity at significant depths have an important influence on the transfer impedance and resulting field from power lines. 12-13

Computer Modeling of Joint Corridors

100,000

Approximate Image Conductor Depth (feet)

10,000

1,000

100 1 10 100 Earth Resistivity (ohm-m) 1000 10000

Figure 12-7 Approximate Image Depth for Earth Return Conductor

Three-Phase Systems The above has only considered a single power conductor. An actual power transmission or distribution line has several conductors. We are primarily concerned about three-phase power transmission and distribution lines, so the principal power conductors are in groups of three. Usually single-circuit (three phase conductors) or double-circuit (six phase conductors) lines are encountered. These power circuits may have one or more shield conductors or neutrals in addition to the phase conductors. Typically, all the conductors carry current and must be included in the calculations to determine the excitation of a nearby rail system. Equation 12-1 provides an expression for the longitudinal electric field for a single current carrying conductor in terms of the conductor current I and the transfer impedance Zt. When more than one current-carrying conductor is present, the resulting longitudinal electric field at a point is the superposition (or sum) of the fields due to each conductor. Thus, the longitudinal electric field due to a three-phase conductor system can be written as:

ET = Z A I A + Z B I B + Z C I C

Eq. 12-3

In Equation 12-3 IA, IB and IC are the currents in the three phase conductors, which are nominally displaced in phase from one another by 120 degrees. Also in the above equation, ZA is the mutual impedance between the observation point and the conductor carrying IA. ZB and ZC are similarly defined for the other phase conductors. Additional components can be added to Equation 12-3 to include the contribution to the field that is caused by the current in a second power circuit or from the shield conductors or a neutral associated with the power line that also conducts current.

12-14

Computer Modeling of Joint Corridors

Therefore the resultant field and the excitation of the railroad conductors is a complex summation of individual components that are related to the current in each conductor of the power system, and is difficult to determine without the use of computers. However, using some simplifying assumptions, we can illustrate some important features of the exciting fields that are useful to keep in mind when preparing or conducting a compatibility study. Many power transmission lines carry current that is nominally the same magnitude in each phase, with the three phase currents displaced by approximately 120 degrees from each other. That is, the line is nominally balanced. Balanced power lines have a couple of interesting properties. For one, the contribution of the image conductors to the total longitudinal electric field becomes negligible. Thus, the current-dependent power-line field that excites the railsystem conductors is not very dependent on the earth resistivity. The second property of balanced power lines is that the field is dependent on the separation of the phase conductors, for observation distances that are not too large. Consider the simple conductor arrangement that is shown in Figure 12-8. Three power conductors A, B, and C are at distances dA, dB, and dC from the observation point q. The magnitude of the currents are approximately equal and their phases are spaced (electrically) approximately 120 degrees apart. Using Equation 12-1 and Equation 12-3, for which D in Equation 12-1 is replaced by dA, dB, or dC, for the individual conductor to observation point distances results in

ET = KI A ln d A + ln d B 120 D + ln d C 120 D

Eq. 12-4

Figure 12-8 Distances from Power Conductors to Observation Point

12-15

Computer Modeling of Joint Corridors

Figure 12-9 shows the phasor components of the total electric field from Equation 12-4 with representative distance values. The three components are 120-degrees out of phase, and would add up to zero if it were not for the separation between conductors, which results in the distances dA, dB, and dC being not the same. As the observation point becomes farther from the phase currents, the distance from each phase to the observation point become more the nearly the same as the distance from the other phases. That is, as the distance q becomes large, dA dB dC. For that condition, the bracketed term in Equation 12-4 tends to zero.

Figure 12-9 Phasor Components of Total Longitudinal Electric Field (LEF) for Figure 12-8

Since higher voltage-class power lines have larger spacing between the phase conductors, the magnetic field intensity of higher voltage lines will usually be higher for the same amount of load current. The fields will be higher not directly because of the voltage is higher, but because the conductors are spaced farther apart. This is also the reason why the fields from a buried power line are less than for an above-ground line (for similar spacing to the observation point), since the phase conductors are closely spaced for buried power lines. The phasor diagram of Figure 12-9 helps to provide some understanding of how the longitudinal electric field of a balanced three-phase transmission or distribution line is related to the geometry of the line, particularly the phase-to-observation point distance. However the approach is difficult to relate to specific values because the math is too complex for easy interpretation. That is the reason that almost all coupling investigations are performed using computer modeling of the geometry of the power line and its relation to the location of the railroad conductors.

12-16

Computer Modeling of Joint Corridors

Figure 12-10 shows the calculated total longitudinal electric field for a horizontal-configuration transmission line, similar to the arrangement of Figure 12-8. The field strength is plotted against the distance from the center of the power line. The distance has been normalized with respect to the phase conductor spacing, s. Different curves are shown for different mean height, h, of the phase conductors, also normalized with respect to the phase-conductor spacing. The electric field units are volts per mile (volts/mi), and that is normalized by the line current in thousands of amperes (kA). Thus the field is shown on the figure vertical axis as volts/mi/kA. As an example, for a distance from the line of 10s (10 times the phase spacing) the field is approximately 20 volts per mile for 1000 amperes of balanced current in the line. That field would result in approximately 10 volts rail-to-ground for a one-mile track circuit with high-resistivity ballast.
1000

M a g n itu d e o f B a la n c e d C o m p o n e n t o f L E F , E (V /m i/k A )

100

h/s = 1.0 1.5 2.0 2.5


10

3.0

1 0.1 1 10 100

Normalized Distance (d/s)

Figure 12-10 Normalized Balanced Component of LEF for Single Circuit Horizontal Line with No Shield Wires

For comparison, Figure 12-11 shows the calculated total longitudinal electric field for a vertical-configuration transmission line. That is, the phase conductors are placed one above the other. For distances that are farther than approximately twice the phase spacing, as measured from directly below the phase conductors, comparison of Figure 12-10 and Figure 12-11 illustrates that the vertical line has a lower field than a similarly spaced horizontal line.

12-17

Computer Modeling of Joint Corridors

1000

Magnitude of Balanced Component of LEF, E0 (V/mi/kA)

h/s = 1.5 100 h/s = 2.375 h/s = 4.125 h/s = 3.25 h/s = 5.0

soil resistivity = 1000-ohmm

10

0.1 0.1 1 10 100

Normalized Distance (d/s)

Figure 12-11 Magnitude of Balanced Component of LEF for Single Circuit Vertical Circuit without Shield Wires

When the magnitude of the currents in all three phase conductors are not all equal, or the phases of the three currents are not ideally displaced in phase from each other (120 degrees), an unbalanced current will flow on the power line. The unbalance current or residual current is an earth return current that is divided among the three phase conductors and flows to ground at the grounded connection of Y-connected transformers. The field components from the residual current components in the three phase conductors tend to be in-phase with each other and thus tend to add. Therefore, the LEF that results from power line residual current is much higher per unit of current than for the balanced component of the current. An unbalance current that is a few percent of the balanced current can produce a longitudinal electric field that is comparable in magnitude to that of the balanced component of the field. Figure 12-12 shows a plot of the unbalanced component of LEF for the vertical line used for the balanced component in Figure 12-11. The field units are in volts/mile and the field is normalized by the magnitude of the residual current in amperes. Thus, to calculate the unbalanced component of LEF in volts/mile, multiply the vertical scale reading by the residual current in amperes. It is seen that the field falls more slowly with distance for the unbalanced component of LEF than for the balanced component.

12-18

Computer Modeling of Joint Corridors

10

Magnitude of Unbalanced Component of LEF, E0 (V/mi/A)

= earth resistivity = 1000 ohmm

h/s = 1.5

h/s = 5.0 (0.875 increments)

1 0.1 1 10 100

Normalized Distance (d/s)

Figure 12-12 Normalized Residual Component of LEF for Single Vertical Circuit with No Shield Wires

Unbalance Comparison Example It is informative to compare the value of residual current that will produce approximately the same value of LEF as the balanced component. For this example, we can use Figure 12-11 for the balanced condition and Figure 12-12 for the unbalanced condition. Assume a vertical line with 25-ft phase spacing, i.e. s = 25 ft, and a 50-ft mean height for the lowest phase, i.e. h/s = 2. Assume a distance of interest of 100 feet from the line, or d = 4s. To calculate the balanced component of LEF, refer to Figure 12-11 where the LEF is approximately 22 V/mi at d = 4s and h/s = 2, for 1000 amperes balanced line current. Assume for simplicity that the soil resistivity is 1000 ohmm, so that the assumptions for Figure 12-12 apply. At d = 4s and h/s = 2, the unbalanced component of LEF is approximately 1.95 V/mile per amp of residual current. In order for the unbalanced component of LEF to be approximately the same as the balanced component, the residual current would have to be: (22 V/mi)/(1.95 V/mi) = 11 A. Thus, if the residual current is 11 amperes, the unbalanced component of LEF is approximately the same as the component of LEF that is caused by 1000 amperes of balanced current. Therefore for this example, a 1.1% unbalance will result in approximately equal contribution to the LEF by balanced and unbalanced components of power-line current. Lower values of soil 12-19

Computer Modeling of Joint Corridors

resistivity will result in lower values of unbalanced-current transfer function. Therefore, for lower values of soil resistivity, a higher percentage of unbalance can be tolerated to cause comparable contribution to the total LEF by the balanced and unbalanced transmission line currents. More on Power Line Unbalance As illustrated in the previous example, rather small percentages of steady state residual (unbalanced) current on a power line can be important in trying to assess what voltages may be induced onto a nearby rail system. Steady state unbalance tends to be ignored by the power industry, since it is in general relatively low, and does not cause undue problems for the power system. Residual current in the transmission line can be caused by the magnitude of the current being different in the three phases, or the power factor being different for the three phases, or any combination of the two. For long transmission lines, some unbalance may result from asymmetric coupling between the conductors or to earth. Considerations of steady-state unbalance is often ignored in rail-system coordination studies, and definitive data on transmission and distribution line unbalance is not readily available in the literature. Recent measurements of unbalance on many power lines of one electric utility characterized the unbalance of several voltage classes, for many lines of each voltage class, using phase current values recorded by the SCADA system [79]. The use of phase current data to estimate threephase line unbalance is not very accurate, because the phase relationships between the currents are not properly included. A much better procedure is to measure the residual current directly. However, that may not always be possible for individual lines. Figure 12-13 shows a plot of the approximate effective residual current that we derived from the data presented in Reference [79]. A trend line is shown as a best fit to the measured data. The data shows that the residual current decreases as the line voltage becomes higher. Although the results of Figure 12-13 may not be representative of all power systems, it is representative of one major system, and similar data is not known to be available for other systems. There does not seem to be consensus on how to account for steady-state power-system unbalance for compatibility analysis. Often there is not a good guideline for selecting the percent unbalance to use for a compatibility study. The information presented in Figure 12-13 may provide some guidance, but most utilities will estimate that their system has significantly less unbalance than is suggested by Figure 12-13. Most utilities seem to be relatively comfortable with assuming 3% residual for most transmission lines up to 345-kV. There is a tendency to model the phase conductor currents at 120-degree phase displacement between the three phase currents, with one or two of the phase current magnitudes changed in magnitude by the percentage unbalance desired for the analysis. This procedure forces the magnitude of the unbalance to be some value and the phase of the unbalance current to be at a fixed (unknown) value with respect to the balanced (positive-sequence) current.

12-20

Computer Modeling of Joint Corridors


100

Residual Current, % of Average Line Current

Average Measured Unbalance for Each Voltage Class, Cramer (4), Converted to Equivalent Residual Current Values

10

Trendline:

Residual(%) = 39.91V -0.3653

1 1 10 Line-Line Voltage (kV) 100 1000

Figure 12-13 Measured Phase Current Deviation from the Circuit Mean

Since in general the phase relationship between the balanced (positive sequence) current and the unbalanced current (residual current) is unknown, thus the phase relationship between the balanced and unbalanced components of the LEF is also unknown. However, the two components of LEF are phasor quantities and will be present at the same time. Therefore, we feel that the safe approach is to separately calculate the balanced and the unbalanced components of the LEF, or the balanced and unbalanced components of rail-induced voltages and currents. Since the two quantities are phasors with unknown phase, they might have almost any phase difference, and the magnitude of the resultant phasor summation could range anywhere from the sum to the difference between the individual magnitudes. A conservative approach is to separately determine the balanced and the unbalanced components of LEF, or induced voltage or current, then algebraically add the two components together, as though they were in-phase. The following paragraphs illustrate an example of such a procedure. Conservative Effect of Steady State Unbalance Example A planned 138-kV single-circuit vertical-configuration transmission line parallels an existing single-track rail line. The exposure length is approximately six-miles. The separation between the centerlines of the transmission line and the track is relatively constant at 35 feet. The phase conductors are offset from the pole by approximately 7 feet on the side of the track, and are spaced by 14 feet. A single 7#7 Alumoweld multi-grounded OHSW is planned at approximately 14 feet above the top phase. The expected full load transmission line current is approximately 1800 amperes, with a maximum of 3% residual current. The nominal soil resistivity is 100 ohmm.

12-21

Computer Modeling of Joint Corridors

Figure 12-14 shows the predicted rail-to-ground voltage profile along the track within the exposure region. Three curves are shown in the figure, the lowest-value curve is the rail-toground voltage caused by just the 3% residual current. The next higher curve is the result of excitation by only the balanced current. The uppermost curve is the algebraic summation of the other two curves. Since the phase relationship between the balanced (positive-sequence) and unbalanced (residual current) is unknown, the upper curve represents the worst-case summation of the contributions from the expected balanced and unbalanced currents. The actual induced voltage could be any value between the upper curve and the middle curve minus the lower curve (not shown), depending on the phase between the balanced and unbalanced components of transmission line current.
160.0

140.0

120.0
Balanced + 3 % un-balance

Rail-to-Ground Voltage (Volts)

100.0
Balanced

80.0

60.0

40.0
3 % un-balance

20.0

0.0 194.0 -20.0

196.0

198.0

200.0

202.0

204.0

206.0

-40.0

Railroad Mile Post

Figure 12-14 Predicted Rail-to-Ground Voltage Including 3% Current Unbalance

Double Circuit Arrangements The foregoing has reviewed several basic forms of single circuit power arrangements. The longitudinal electric field that is created by a power line becomes considerably more complex when two power circuits are operated together on the same structure. There are many different possible arrangements of the phase relationships for the conductors of two power circuits. Figure 12-15 summarizes several possible arrangements for the phasing of two vertical circuits. The most important two arrangements for EMI considerations are the top two phase assignments shown in the figure. The first phasing arrangement of Figure 12-15 results in the lowest LEF of all possible phasings, if equal current loading exists on the two circuits. That is because at any observation point, the field from one circuit tends to cancel the field from the other circuit. This advantage deteriorates when the current in the two circuits is not the same, or 12-22

Computer Modeling of Joint Corridors

if unbalanced current exists in one or both circuits. Figure 12-16 shows the normalized longitudinal electric field for the center point symmetric phasing of a double-circuit vertical arrangement. Comparison with the curves of Figure 12-11 for a single vertical circuit shows that the double circuit, with twice the current (that is, the field is normalized to 1kA per circuit), has lower longitudinal electric field, by a factor of four or so, depending on the location.

Figure 12-15 Summary of Vertical Double Circuit Phase Assignments

12-23

Computer Modeling of Joint Corridors


Circuits are separated by 1.5s 1000

Magnitude of Balanced Component of LEF, E(V/mi/kA)

100

h/s = 1.5

s is spacing between phases (feet). f is mean height of bottom phase. d is distance to observation point from structure. both circuits with equal load.

10 h/s = 5.0 (0.875 increments)

0.1

0.01 0.1 1 10 100

Normalized Distance from Center of Power Circuit (d/s)

Figure 12-16 Normalized Balanced Component of LEF for Double Circuit Vertical Center Point Symmetric Line with No Shield Wires

The center point symmetric phasing arrangement is the preferred phasing for a vertical double circuit power line. However, as has been noted, the cancellation benefits may not be achievable on a continual basis because of the variable loading of the two circuits. In fact, the worst-case condition, which must be considered, is when one of the circuits is out of service. When one of the circuits is out of service, the loading on the circuit that remains in service may increase above the normal full load for the circuit when both circuits are in service. While the benefits of the center point symmetric phasing arrangement may not always be in effect, most of the time a reduced-field benefit will be achieved. There have been systems built that have forced equal current to flow on both circuits as a mitigation measure, by tying the two circuits in parallel in the center point symmetric configuration. Many older double circuits are phased as is shown in the second arrangement shown in Figure 12-15. That super-bundle arrangement is the worst phasing arrangement of all for a doublecircuit power line. The fields from that arrangement are higher than for any of the other phasings. As is shown in Figure 12-17, the LEF for this circuit arrangement is considerably higher than for the center point symmetric arrangement of Figure 12-16. Where possible, the second phasing arrangement of Figure 12-15 should be avoided.

12-24

Computer Modeling of Joint Corridors


Circuits are separated by 1.5s

1000

Magnitude of Balanced Component of LEF, E(V/mi/kA)

h/s = 1.5 100 h/s = 5.0 (0.875 increments)

10

s is spacing between phases (feet). f is mean height of bottom phase. d is distance to observation point from structure. both circuits with equal load.

0.1 0.1 1 10 100

Normalized Distance (d/s)

Figure 12-17 Normalized Balanced Component of LEF for Double Circuit Vertical Super Bundle with No Shield Wires

Steady State Compatibility Modeling of Railroad Systems for Magnetic Field Coupling from Power Systems The foregoing section has described how the power line currents result in a longitudinal electric field at the location of railroad conductors such as a rail or other signal or communications conductors. The following section will show how that exciting field causes a voltage to be developed on the rails and how some of the properties of the track system and the soil influence the resulting voltage that is induced onto the track. Track Signal Circuits and Their Effect on Induced Voltage Railroad systems that have fixed-block train-control signaling have segments of track that are separated by insulating joints. See Chapter 4 for more comprehensive discussion of railroad signaling and track circuits. The length of track between sets of insulated joints is a track circuit. Train-control equipment is connected rail-to-rail at each end of a track circuit. Numerous types of equipment are used for train control, but a common characteristic is that a signal is transmitted from one end of the track circuit to the other using the two rails as two-conductor signal circuit. A signal source applies the signal rail-to-rail at one end of the circuit, and a signal receptor senses the presence of the rail-to-rail signal at the other end of the signal circuit. The traincontrol signal systems are generally dc or very-low-frequency pulsed systems. Other signal

12-25

Computer Modeling of Joint Corridors

systems, for example for highway grade crossing control, may also exist within a track signal circuit, or overlap into more than one basic track circuit. An important factor is that the railroad signal system components of concern are connected between the two rails, with the two connections relatively close together, in terms of the overall length of the signal circuit, such that for most practical purposes the connection points can be considered to be at the same location on each rail. Figure 12-18 is a simple sketch of a single track signal circuit. Generally, additional track circuits will exist to the right and left of the one shown. These other track circuits are isolated from the circuit that is shown by insulated joints and are therefore not shown.

Figure 12-18 Longitudinal Electric Field Excitation of Rails and Induced Voltages

The figure depicts the longitudinal electric field excitation of the two rails as E1 and E2 at the arbitrarily labeled rail #1 and rail #2. Also labeled are the rail-to-ground voltages and the rail-torail voltages at the two ends of the track circuit. The rail-end to ground voltages are typically the highest voltage within a track circuit, and therefore are of concern for personnel safety. The sketch shows impedances z1 and z2 that represent the input impedance to the track signal equipment at the two ends of the track circuit. The rail-to-rail voltage at the ends of the circuit, across the signal equipment is important, since those voltages can cause interference to the operation of the train-control signal equipment, if the voltage is sufficiently high. Rail-to-rail voltage at other locations may also be important if signal equipment, such as highway grade crossing equipment, is connected at those locations. The longitudinal electric field excitation shown in Figure 12-18 results in an equivalent incremental voltage source being induced into the rail, all along its length. Figure 12-19 shows a schematic of a typical incremental (small) segment of a long section of rail that is excited by a longitudinal electric field. A rail of a track circuit that is excited by the magnetic field of a 12-26

Computer Modeling of Joint Corridors

transmission line is comprised of many such incremental segments connected one-to-another. Three elements are shown in the schematic, these are:

E dx is an equivalent voltage source, where E is the component of the exciting longitudinal electric field of the power line at the location of the rail segment. The units of E are (volts per unit length) and the unit length is generally a meter, but can be expressed as any other unit of length, such as per kilometer or per mile. The quantity dx is the incremental length of the equivalent circuit. To develop the differential equations that describe the conductor current, dx becomes infinitesimally small, but can be thought of as being the same per unit length as the longitudinal electric field, such that the product E dx is just a voltage source that is valid for the rail segment in the schematic. Z dx is the equivalent series impedance of the rail segment. Z is the series impedance per unit length of the rail, which includes both the internal and external (as influenced by the earth) impedance of the rail conductor, and as before, dx is the incremental length of the equivalent circuit. Y dx is the equivalent admittance (reciprocal of impedance) of the rail segment. Y is the admittance per unit length, which accounts for the leakage of current from the rail through the resistivity of the ballast, and includes the admittance resulting from the flow of current from the rail into the earth. Again, dx is the incremental length of the equivalent circuit.

Figure 12-19 Equivalent Circuit of Incremental Segment of Rail

The schematic shows that the current entering the left side of the rail segment can be different than the current that flows out the right side of the segment, which accounts for the current that enters or leaves the rail over the length of the segment, through the admittance to earth. Similarly, the schematic shows that the voltage at the left side of the rail segment can be different than at the right side as a result of the voltage source E dx and the voltage drop (I Z dx) across the series impedance Z dx. 12-27

Computer Modeling of Joint Corridors

Including additional conductors, such as other rails or other conductors, makes the equivalent circuit more complex, since mutual impedance and admittance elements between the various conductors must be accounted for. The mutual impedances account for the influence that current in one conductor has on another conductor. Actually, the mutual impedance modifies the longitudinal electric field component E to account for the current flow in other conductors. Also, current that flows to or from one conductor within the equivalent circuit segment length can cause a voltage at the location of other conductors, which in turn can result in additional current entering or leaving the other conductor. That coupling mechanism results in admittance elements Y that connect between different conductors. The computer programs that have been developed to analyze power system coupling to railroad systems account for the coupling of voltage and current from the power system conductors to the railroad conductors, and they also account for the coupling that occurs between passive conductors, such as nearby rails, other signal conductors, as well as non-railroad system conductors such as nearby pipelines. The voltage that will appear from rail to ground at the end of a track circuit, as a result of power line excitation and the connection together of numerous equivalent-length segments such as in Figure 12-19, depends on:

The values of the three elements of the equivalent circuit, How the equivalent circuit elements vary along the length of the track circuit.

Most of the time, information on existing or proposed common corridors is quite sketchy. Generally the element of the equivalent circuit that is most readily determined is the longitudinal electric field, since it is a function of:

The geometry of the power line, The spacing between the power line and the railroad conductors, The soil resistivity, The current in the power line phase conductors, The balance of the current in the phase conductors, The current in the OHSW or the neutral conductors of the power line.

Some of these quantities may not be known very accurately. However, for a proposed transmission line, more-or-less worst-case assumptions can be made to develop predicted railsystem voltage profiles that are a conservative estimate of the voltages that should be anticipated. A more difficult task is to compare measured values of, for example, rail-to-ground voltage with predicted values for an operating power line, since it is quite difficult to obtain adequate characterization of all the power and rail-system parameters to permit a valid comparison of measured and predicted values. As noted, the voltage that is induced along the length of a railroad track circuit by magnetic-field coupling from a power line results from stringing together a group of equivalent circuits, such as Figure 12-19, for both rails over the length of the track circuit. The following discussion addresses how some of the parameters in the equivalent circuit affect the induced rail-to-ground voltage. 12-28

Computer Modeling of Joint Corridors

The primary factors that influence the series impedance in Figure 12-19 include the frequency of the exciting waveform, the size of the rail and its electrical properties, and the resistivity of the earth. For a fixed frequency and typical rail properties, variation of soil resistivity over a range from 10-to-1000-ohmm causes the impedance of the rail to change by less than 20%. Thus, the primary element that can change in the equivalent circuit of Figure 12-19 is the admittance, which depends on the track ballast resistivity and the soil resistivity. The following three figures (Figure 12-20 through Figure 12-22) illustrate how the predicted rail-to-ground voltage at the end of a track circuit (normalized by the exciting longitudinal electric field) varies as a function of the length of the track circuit, the soil resistivity and the ballast resistivity. Each figure is for a different value of soil resistivity (ohmm). Within each figure, the calculated track circuit rail-to-ground voltage, as a function of track-circuit length is shown for several values of ballast resistivity, ranging from a low of 3 ohmkft to 100 ohmkft. Figure 12-20 is for 10-ohmm soil and shows that for high ballast resistance conditions, the induced voltage at the ends of the track circuit increases as the length of the track circuit increases. However, for low (3-ohmkft) ballast, track circuits that are longer than approximately one-mile do not have higher voltage. Thus, for most soil resistivity and ballast resistivity values, the voltage at the ends of a track circuit that is induced by a uniform magnetic field of a power line, tends to increase as the length of the track circuit increases.

1.50

Parameters Soil Resistivity: 10 Ohms Zbal is Rail-to-Rail Ballast Leakage Resistance Per 1000ft Track

Zbal = 300ohms 100ohms 50ohms 30ohms

1.25

Rail Voltage / Electric Field (Volts/Mile)

1.00

0.75

10ohms

0.50 3ohms

0.25

0.00 0.0 0.5 1.0 1.5 2.0 2.5 3.0

Track Block Length (Miles)

Figure 12-20 Normalized Voltages, End of Rail-to-Remote Earth

12-29

Computer Modeling of Joint Corridors


1.50

1.25

Parameters Soil Resistivity: 100 Ohmm Zbal is Rail-to-Rail Ballast Leakage Resistance Per 1000ft Track

Zballast = 300ohms 100ohms 50ohms 30ohms

Rail Voltage / Electric Field (Volts/Mile)

1.00

10ohms 0.75

3ohms 0.50

0.25

0.00 0.0 0.5 1.0 1.5 2.0 2.5 3.0

Track Block Length (Miles)

Figure 12-21 Normalized Voltages, End of Rail-to-Remote Earth

1.50

Parameters Soil Resistivity: 1000 Ohmm Zbal is Rail-to-Rail Ballast Leakage Resistance Per 1000ft Track

1.25

Rail Voltage / Electric Field (Volts/Mile)

Zbal = 300ohms 100ohms 50ohms 30ohms 10ohms 3ohms

1.00

0.75

0.50

0.25

0.00 0.0 0.5 1.0 1.5 2.0 2.5 3.0

Track Block Length (Miles)

Figure 12-22 Normalized Voltages, End of Rail-to-Remote Earth

12-30

Computer Modeling of Joint Corridors

Rail ballast resistivity can vary over a significant range as a function of whether conditions for a given exposure. Ballast resistivity can also vary significantly along the length of an exposure as a result of ballast maintenance, chemical spills or loss of material from railcars, or salt spreading at highway crossings in the northern climates. Tests that were made during a controlled EPRI/AAR field test illustrate the effect of weather on the short-term resistivity of ballast. The tests were conducted in October 1991 in Ohio to corroborate and track signal equipment tests in a laboratory interference simulator. The test period included both hot and dry as well as periods of heavy rain. Two half-mile segments of Conrail track were dedicated for the testing. Both ac and dc measurements of ballast resistivity were made, using the circuit of Figure 12-23. The sketch shows the location of voltage measurements that were obtained. The sketch shows a battery and its current limiting resistor, Rb, and a 0.01 ohm resistance for measuring the current. Rail-to-rail and rail-to-ground voltage measurements were made on each end of the track segment. Measurements were also made with an ac signal source in place of the battery.

Figure 12-23 Ballast Resistivity Test Arrangement

A data sheet for recording the measured data and calculated ballast resistivity is shown as Figure 12-24. The measured ballast resistivity for both ac and dc excitation of the track signal circuit is shown in Figure 12-25. Ballast resistivity data were obtained separately on two half-mile segments of track circuit, the east and west half of a one-mile track signal circuit.

12-31

Computer Modeling of Joint Corridors

Ballast Measurement Data Sheet, RP 1902-6

Date _________ W mile ?

Time _________ E mile ?

d.c. ?

a.c. ?

One mile ?

Z (kft) = _______________

Composite Ballast Measure VRRb = _______________ VRRa = _______________ Ib = VRRb / 0.01 = _______________ Calculate (ohm*kft) RBal = Z(VRRa + VRRb) / 2Ib = _______________

Individual Ballast Measure VNRa = _______________ VSRa = _______________ VNRa = _______________ VSRa = _______________ Note: Calculate (ohm*kft) RNBal = Z(VNRa + VNRb) / Ib = _______________ RSBal = Z(VSRa + VSRb) / Ib = _______________

Figure 12-24 Ballast Measurement Data Sheet

12-32

Computer Modeling of Joint Corridors


50

AC East DC East AC West DC West

40

Ballast Resistivity (ohm*kft)

30

20

10

0 9-30/1600 10-3/1700 10-4/1000 10-4/1500

Measured Date/Time

Figure 12-25 Measured Ballast Resistivity at Talmadge, Ohio

Figure 12-25 illustrates the results of the ballast testing in which the horizontal scale displays the date and time of the measurements. The ballast resistivity for Monday, September 30 is typical for a dry condition, as no rain had occurred for several hot days. Rain occurred on Wednesday; thus the ballast resistivity was lower on Thursday, October 3. Hard rain fell during the morning of Friday, October 4. The effect of that rainfall on the ballast resistivity evident for the first group of data on October 4. The last ballast-resistivity measurements were made in the afternoon of Friday, October 4, after the sun had been out for a few hours. The effect of the drying of the ballast can be seen in the last measurements of October 4 that are shown in Figure 12-25. It will be noted in Figure 12-25 that the ac ballast resistivity is generally less than the dc ballast resistivity. One reason for this can be attributed to what is called the track storage effect. The track storage effect results from the electrochemical double-layer capacitance at the surface of the rail in contact with an electrolyte. Figure 12-26 shows an equivalent circuit of a metalelectrolyte boundary. The resistance Rp is the polarization resistance, which is a nonlinear function of the current through the metal-electrolyte boundary. The capacitor that is shown in the figure is the double-layer capacitance. This capacitance can be quite large, in the range of 10-1000 microfarad per square centimeter, for a metal in direct contact with a high moisturecontent electrolyte. The other resistance shown in the equivalent circuit of Figure 12-26 is a spreading resistance Rs, from the metal-electrolyte interface to remote earth, through the ballast resistivity and the underlying soil.

12-33

Computer Modeling of Joint Corridors

Figure 12-26 Equivalent Circuit at Metal-Electrolyte Interface

The effective value of the capacitive component of the ballast leakage is a function of the moisture present and not likely to be a constant with time or location along a rail. Direct current flowing from the rail must flow through both resistances of the equivalent circuit. Alternating current tends to bypass the polarization resistance and flows through the capacitance. As the capacitance is very large, even quite low frequencies tend to bypass Rp. Common experience indicates that the 60-Hz ac-ballast resistance is often in the range of 60 to 70 percent of the dc-ballast resistance. That range tends to be consistent with the results that are shown in Figure 12-25. Track Unbalances and Induced Voltage at Signal Equipment An important element of inductive coordination is to identify if existing railroad signal equipment is compatible with the voltage that is induced onto the railroad signal conductors. Typically, personnel safety guidelines may be easier to meet than signal-equipment compatibility issues. Therefore, control of personnel safety issues is generally addressed first, since often the tools available for controlling personnel safety are different than for control of signal system compatibility. The methods used to control personnel safety will also generally help reduce interference to the signal systems, but the converse may not be true. Often the approach to reducing personnel safety issues is to reduce the exciting field at the rail-system conductors. Specific methods for reducing the exciting field at the rail-system conductors will be reviewed in Chapter 13. Generally those methods will reduce the coupled power-system interference voltage that is developed at the inputs to the signal equipment. However, the approaches that are used to reduce the effective exciting fields are often not adequate to ensure the compatibility of the power system with the signal system components. 12-34

Computer Modeling of Joint Corridors

The railroad-system conductor voltages that are important for personnel safety considerations are generally the voltages between the conductors and earth. Often the conductors of interest are the rails. As was discussed earlier in this chapter, for a given uniform field excitation, the induced rail-to-earth voltage depends on such factors as the track-circuit length, the ballast resistivity, and the soil resistivity. The voltage across signal-equipment is also dependent on the rail-to-earth voltage. However, since the signal equipment of interest is almost always connected between two signal-system conductors the voltage that is of concern is the voltage between the two conductors. The conductor-to-conductor voltage, for example between two rails, is typically in the range of a few percent of the induced rail-to-ground voltage for normal track conditions, but can be considerably higher for some unbalancing conditions. Important unbalancing conditions should be evaluated in a compatibility investigation and for mitigation design. Unbalancing conditions may include a shorted insulated joint, an earthed lead-in conductor (with or without a train within the affected track region), or a locomotive within the stagger region at an insulated-joint location. Figure 12-27 illustrates the measured insulated-joint resistances for an approximately 13-mile segment of track, using a commercial IJ resistance test instrument. The average IJ resistance is 40.4 ohms. While the resistance varies over approximately a 3:1 range. It is seen that the unbalance of joint-pairs is relatively small. Thus for this measurement set, the small existing unbalance may not result in interfering conditions, depending on the induced rail-to-ground voltage and the types of signal equipment.

Figure 12-27 Measured Resistance of Rail Insulated Joints

12-35

Computer Modeling of Joint Corridors

In view of the above measured results that illustrate that the resistance of rail insulated joints can be in the tens-of-ohms range, it is of interest to review how the resistance of insulated joints can influence the voltage that is induced onto the rails by the fields of a power line. As an example, consider several track circuits that are uniformly excited by the field of a power line. For the example, we have assumed that the power line joins the railroad ROW and parallels it for the distance of six track circuits before departing, with each track circuit assumed to be 1.5 miles long. If all the insulated joints are assumed to be nominally the same resistance, the rail-to-ground voltage depends of the value of IJ resistance, and other parameters, such as soil and ballast resistivity. The three following figures show the calculated induced rail-to-ground voltage at IJ locations within a hypothetical exposure region that has an assumed constant LEF excitation, soil resistivity (100 ohmm), and track ballast (100 ohmkft), for different assumed value of IJ resistance. For each figure, the rail-to-ground voltage is normalized by the voltage if all IJs were a very high (1000-ohm) resistance. Figure 12-28 shows the calculated voltages with all IJ resistances at 70 ohms. Rail-to-ground voltages on the north side of the IJs are shown as lightcolored bars, and dark bars represent the voltage on the south side of the IJ. It is seen the voltages only vary by approximately 10% over the exposure.

1.6

North
1.4

South of IJ

North of IJ

1.2

Normalized Rail-Ground Voltage

1.0

100 ohmm soil resistivity. 100 ohm kft ballast resistivity. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All IJ's have a 70-ohm resistance.

0.8

0.6

0.4

0.2

0.0 1 2 3 4 5 6 7

IJ Locations for 1.5-mile Track Circuits.

Figure 12-28 Calculated Rail-to-Ground Voltage at IJ Locations for All 70-ohm IJs

12-36

Computer Modeling of Joint Corridors

Figure 12-29 shows the calculated rail-to-ground voltage at IJ locations for 30-ohm resistance IJs. It is seen that the track circuits near the ends of the exposure have IJ-voltages that differ by approximately 20% from the values expected for high-resistance joints. The effect of all 10-ohm insulated joint resistance values is illustrated in Figure 12-30. The low-resistance IJs used to develop this figure result in significant deviation in rail-induced voltage from the value to be expected for high-resistance joints. It is unlikely that all insulated rail joints will the same resistance, as is the case for this example, however, the data of Figure 12-27 shows that the resistance of many IJs tend to fall into a range similar to the value used for Figure 12-29.
1.6

1.4

North

South of IJ

North of IJ

Normalized Rail-Ground Voltage

1.2

1.0

Soil resistivity 100 ohm m. Ballast resistivity 100 ohm kft. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All IJ's have 30-ohm resistance.

0.8

0.6

0.4

0.2

0.0 1 2 3 4 5 6 7

IJ Locations for 1.5-mile Track Circuits.

Figure 12-29 Calculated Rail-to-Ground Voltage at IJ Locations for All 30-ohm IJs

These considerations may be particularly relevant when attempting to compare predicted rail voltages against measured values to evaluate the validity of predictions. Generally the values of IJ resistance are unknown during a predictive study, so it should not be surprising if some deviation between measured and predicted rail voltage occurs. One cause for such discrepancy may be the unknown values of the IJ resistances. The above figures also show that the highest values of IJ resistance do not necessarily result in the highest values of induced rail voltage. As the IJ resistance changes uniformly, these figures show that the voltages at some locations increase, while the voltage at other locations decrease. However, the average voltage remains approximately constant. The Figures 12-28 through 12-30 show that lower values of (uniform) insulated-joint resistance tend to result in higher variation of rail voltage at IJ locations within an exposure.

12-37

Computer Modeling of Joint Corridors

1.4

South of IJ

North of IJ

North
1.2

Normalized Rail-Ground Voltage

Soil resistivity 100 ohm m. Ballast resistivity 100 ohm kft. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All IJ's have 10-ohm resistance.

0.8

0.6

0.4

0.2

0 1 2 3 4 5 6 7

IJ Locations for 1.5-mile Track Circuits.

Figure 12-30 Calculated Rail-to-Ground Voltage at IJ Locations for All 10-ohm IJs

The results presented in Figures 12-28 through 12-30 are all for a 100-ohmm soil resistivity. Higher soil resistivity or ballast resistivity causes larger extremes in the voltage variation, whereas a lower value of soil or ballast resistivity reduces the voltage variations. These trends are illustrated in the following figures. Figure 12-31, which has 1000-ohmm soil resistivity along with uniform 10-ohm IJ resistance has greater voltage variation than Figure 12-30 which has similar parameters but, 100-ohmm soil resistivity. An example of the smoothing effect on the rail-voltage that results from low ballast resistivity is illustrated in Figure 12-32, which has 5-ohmkft ballast resistivity, but other parameters the same as for Figure 12-30. Figure 12-32 shows that the voltage is almost the same at all IJ locations within the exposure.

12-38

Computer Modeling of Joint Corridors


1.6

North
1.4

South of IJ

North of IJ

Normalized Rail-Ground Voltage

1.2

1000-ohmm soil resistivity. 100-ohmkft ballast resistivity. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All Ij's have 10-ohm resistance.

0.8

0.6

0.4

0.2

0 1 2 3 4 5 6 7

IJ Locations for 1.5-mile Track Circuits.

Figure 12-31 Calculated Rail-to-Ground Voltage at IJ Locations for 10-ohm IJs


1.6

North
1.4

South of IJ

North of IJ

Normalized Rail-Ground Voltage

1.2

100-ohm-m soil. 5-ohmkft ballast resistivity. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All Ij's have 10-ohm resistance.

0.8

0.6

0.4

0.2

0 1 2 3 4 5 6 7

IJ Locations for 1.5-mile Track Circuits.

Figure 12-32 Calculated Rail-to-Ground Voltage at IJ Locations for 10-ohm IJs

12-39

Computer Modeling of Joint Corridors

The foregoing has showed how different uniform values of IJ resistance can influence the rail voltage that can be induced by a simple uniform exposure by a power line. Also of interest is the effect of lower IJ resistance at one location. The next three figures assume representative constant values of soil and ballast resistivity and a uniform exposure by a power line. The resistance of all insulated joints is assumed to be a high value (1000 ohms), except for one pair. Figure 12-33 shows the rail-to-ground predicted voltages at IJ locations, for the resistance of one pair of joints (at location 0.0) assumed to be 70 ohms. The voltages in Figure 12-33 are normalized by the voltage that would result if all IJ resistances were 1000 ohms. It is seen that when the resistance of one pair of IJs is reduced to 70 ohms, the voltage at the ends of both track circuits that are separated by the lower-resistance set of IJs is affected, but by less than 10%. The change is small because 70 ohms is a relatively high value of IJ-resistance for ac-inductive coordination considerations.
2.0 1.8

North
South of IJ North of IJ

1.6

Normalized Rail-Ground Voltage

1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 -4.5 -3.0 -1.5 0.0 1.5

100-ohmm soil resistivity. 100-ohmkft ballast resistivity. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All Ij's have 100-ohm resistance, except at location 0.0 which has 70ohm IJ's. .

3.0

4.5

IJ Locations - Miles from Reduced-Resistance IJ's.

Figure 12-33 Calculated Rail-to-Ground Voltage at IJ Locations for 70-ohm IJ Resistance at Location 0.0

Figure 12-34 is based on analysis parameters that are the same as for the previous figure, with the exception that the pair of IJs at location 0.0 are assumed to be 30 ohms. It is seen that lowering the resistance of the pair of insulated joints from 70 ohms to 30 ohms results in greater voltage deviation at the IJs on either end of the track circuit that has the low-resistance IJs. The 30-ohm IJs result in approximately 20% voltage deviations, relative to the case of all high-resistance IJs. Similarly, if the resistance of a pair of IJs is as low as 10 ohms, the voltage deviations are even greater, as is shown in Figure 12-35. With a 10-ohm resistance for an IJ-pair, the deviation of induced 60-Hz voltage in the track circuits that are connected to the low-resistance IJ is in the range of 50%, relative to the voltage that would occur for high-resistance joints. 12-40

Computer Modeling of Joint Corridors


2.0

North
1.8

South of IJ
1.6

North of IJ

Normalized Rail-Ground Voltage

1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 -4.5 -3.0 -1.5 0.0 1.5

100-ohmm soil resistivity. 100-ohmkft ballast resistivity. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All Ij's have 100-ohm resistance, except at location 0.0 which has 30ohm IJ's.

3.0

4.5

IJ Locations for 1.5-mile Track Circuits.

Figure 12-34 Calculated Rail-to-Ground Voltage at IJ Locations for 30-ohm IJ Resistance at Location 0.0
2 1.8

North
South of IJ North of IJ

1.6

Normalized Rail-Ground Voltage

1.4 1.2 1 0.8 0.6 0.4 0.2 0 -4.5 -3.0 -1.5 0.0

100-ohmm soil resistivity. 100-ohmkft ballast resistivity. Induced voltage normalized by voltage for condition of all 1000-ohm resistance IJ's. Uniform field excitation between IJ #1 and IJ # 7. All Ij's have 100-ohm resistance, except at location 0.0 which has 10-ohm IJ's. .

1.5

3.0

4.5

IJ Locations - Miles from Reduced Resistance IJ's.

Figure 12-35 Calculated Rail-to-Ground Voltage at IJ Locations for 10-ohm IJ Resistance at Location 0.0

12-41

Computer Modeling of Joint Corridors

Figure 12-36 is a different depiction of the results that were presented in Figures 12-34 through 12-36. Figure 12-36 assumes several 1.5-mile long track circuits excited by a uniform power-line field. The resistance of all IJs except one pair is assumed to be high resistance (1000 ohms). The pair of IJs at zero reference distance is set to different values, ranging from 1,000 ohms to a low of 5 ohms. The plot shows the calculated rail-to-ground profile along one rail. The magnitude of the rail-to-ground voltage has been normalized by the maximum voltage with all IJs at 1,000 ohms. With all IJs at 1,000 ohms, the maximum normalized voltage is 1.0, depending on which side of the IJ the voltage is observed, and the rail-to-ground voltage at the center of the track circuit is zero. As the resistance of the pair of IJs at the zero-reference location becomes smaller, the rail-to-ground voltage on both sides of the IJs at that location becomes less, but the voltage at the opposite ends of the those track circuits becomes higher. For 10-ohm IJs at the reference location, the magnitude of the voltage at the opposite ends of those track circuits becomes approximately a factor of 1.5 times higher than with all IJs at a high (1000-ohm) resistance, which is the same result that was displayed in a different format in Figure 12-35.
2 Resistance of IJ's at Zero-Reference. All others 1000-ohms. 5-ohm 1 10-ohm 20-ohm 30-ohm 0.5 50-ohm 100-ohm 1000-ohm 0

1.5

Normalized Rail-ground Voltage

-0.5

-1 Uniform 1.5-mile Track Circuit Length. -1.5

-2 -4 -3 -2 -1 0 1 2 3 4

Distance (miles)

Figure 12-36 Rail-to-Ground Voltage vs Resistance of IJs at Zero Reference

Degraded IJ and Track Unbalance The foregoing examples on the effects of lower values of IJ-resistance have considered a pair of IJs to be approximately the same resistance, which is perhaps an unlikely condition. A more likely condition is for one IJ to become degraded, or have lower resistance than most other IJs. The degradation of a single IJ can have more serious consequences for railroad-system compatibility with a power system than if a pair of IJs at the same location degrade. 12-42

Computer Modeling of Joint Corridors

Degradation of one IJ tends to result in a different rail-to-ground induced-voltage profile along each of the two rails, which results in a voltage being developed between the two rails. Since the signal equipment is connected from rail-to-rail, any induced voltage that appears from rail-to-rail may be impressed across the input terminals of the signal equipment, depending on the location of the signal equipment relative to the location of the degraded joint, and the length of the track circuit. Figure 12-37 illustrates the effect of a degraded insulated joint in a uniformly excited track with 1.5 mile long track circuits. The rail-to-ground induced voltage is shown for several contiguous track circuits. The analysis assumes that one of the insulated joints at the center of the plot has a low resistance, while all other IJs are assumed to be 100 ohms. The rail-to-ground voltage on one rail is a solid line, while the voltage on the rail with the degraded insulated joint is a dashed line. The magnitude of the voltage is normalized to the maximum voltage at the IJ-locations with all good IJs. The negative voltages in the graph are nearly 180 degrees out of phase with the positive voltages. Therefore, the rail-to-rail voltage is approximately the difference between the two rail voltages. Thus, at the degraded IJ-location, the rail-to-rail voltage is approximately 60% of the rail-to-ground voltage with all good IJs. Since the phase angles of the two rail voltages are not really the same, the actual rail-to-rail voltage is the phasor difference which may not be the same as the difference of the rail voltages shown in Figure 12-37. Nevertheless, Figure 12-37 is a reasonable representation of the rail voltage variation near a single degraded IJ.
2.5 2.0 1.5

Uniform Excitation of Rails by Power Field. 1.5-mile Long Track Circuits. 100-ohmkft Track Ballast. 100-ohmm Soil Resistivity.

Rail-to-Ground Voltage (normalized)

1.0 0.5 0.0 -8 -0.5 -1.0 -1.5 -2.0 -2.5 -6 -4 -2 0 2 4 6 8

Shorted IJ Location

IJ Locations Distance From Degraded IJ Site (miles)

Figure 12-37 Rail-to-Ground Voltage Profile With One Shorted Insulated Joint

12-43

Computer Modeling of Joint Corridors

It is of interest to evaluate how the rail-to-rail voltage is affected by the resistance of a degraded insulated joint. Figure 12-38 shows the calculated rail-to-rail voltage for the same conditions as for Figure 12-37. Two curves are shown in Figure 12-38, one (the dashed line) is the rail-to-rail voltage at the degraded IJ location, while the solid curve is the rail-to-rail voltage at the opposite end of the track circuit, that is, at the next insulated joint (which is assumed to be intact). The rail-to-rail voltage at the degraded IJ location is several percent higher than the voltage at the other end of the same track circuit. Figure 12-38 for a single degraded IJ shows that the rail-torail voltage at IJ locations increases rapidly as the degraded IJ-resistance decreases. The calculated rail-to-rail voltage is displayed as a percent of the rail-to-ground voltage at those same locations with high-resistance IJs. For these examples the predicted rail-to-rail voltage at a failed, low-resistance IJ (0.1-ohm) site, is approximately 65% of the normal rail-to-ground voltage. However, the rail-to-rail voltage is approximately 50% for a 3-ohm degraded IJ, and is approximately 25% for a 10-ohm IJ. The calculations for these figures have assumed representative electronic track circuit impedances connected rail-to-rail on each side of each IJ. If dc battery and relay circuits had been assumed, the rail-to-rail voltages may be different, because of the different impedances of the two track-circuit components. However, the above examples illustrate the range of conversion from rail-to-ground voltage to rail-to-rail voltage that can occur.
1000 1.5-mile track blocks 100-ohm m ballast 100-ohm m soil At Adjacent IJ Location

At Degraded IJ Location

Percent Voltage Conversion

100

10

1 0.1 1 10 100

Degraded IJ Resistance (ohms)

Figure 12-38 Conversion of Rail-Ground Voltage to Rail-Rail Voltage vs. Degraded IJ Resistance for 1.5 mile track circuits

The normalized results that are shown in Figure 12-38 are not strongly dependent on the circuit length, the soil resistivity, or the ballast resistivity. For example, Figure 12-39 shows the results for a similar analysis, with one-mile track circuits, but other parameters the same as for Figure 12-38, which is for 1.5-mile circuits. Comparison of Figure 12-38 and Figure 12-39 shows comparable percentage conversion of rail-to-ground voltage into rail-to-rail voltage. 12-44

Computer Modeling of Joint Corridors


1000
1-mile track blocks 100-ohm m ballast 100-ohm m soil At Adjacent IJ Location

At Degraded IJ Location

Percent Voltage Conversion

100

10

1
0.1 1 Degraded IJ Resistance (ohms) 10 100

Figure 12-39 Conversion of Rail-Ground Voltage to Rail-Rail Voltage vs. Degraded IJ Resistance for One Mile Track Circuits

However, the rail-to-ground voltage with good IJs tends to increase with the circuit length, the soil resistivity, or the ballast resistivity. Therefore, while the rail-to-rail voltage as a percentage of the rail-to-ground voltage is in the range of 60% for both 1-mile track circuits and 1.5-mile track circuits, the rail-to ground voltage will be higher for the longer circuit, and thus, the rail-torail voltage will tend to be higher for the longer track circuit. In a similar manner for example, the rail-to-rail voltage as a percentage of the rail-to-ground voltage is in the range of 60% for both 100-ohmkft ballast or 5-ohmkft ballast. However, the actual rail-to-ground voltage will be higher for 100-ohmkft ballast. Therefore, the rail-to-rail voltage should be higher for 100-ohmkft ballast resistivity than with 5-ohmkft ballast resistivity, for constant other parameters. AREMA qualification testing for insulated joints by the manufacturer (AREMA Specifications 2.11.7) recommends the resistance of fully assembled bonded insulated joint have a minimum resistance of ten (10) megohms. The insulated joints should also tolerate 3000 volts, 60-Hz ac, for not less than 60 seconds. The results of Figure 12-38 or Figure 12-39 are for much smaller insulated joint resistance than for AREMA specifications for a newly manufactured insulated joint. However, there seems to be no good data available on the degradation of in-service insulated-joint resistance, or whether the presence of induced voltage across IJs has any influence on the failure of insulated joints.

12-45

Computer Modeling of Joint Corridors

Figure 12-38 and Figure 12-39 indicate that an insulated joint does not have to be really bad to result in a significant unbalance of the track for power line induced voltage. Typical track signal circuits will operate satisfactorily with one or more insulated joints less than 10 ohms. Often, the signal system will operate satisfactorily with at least one very-low resistance insulated joint and the signal system performance is not significantly affected (in the absence of interference) until a second insulated joint fails. However, these figures show that if significant rail-to-ground induced voltage is present, a degraded insulated joint may cause a large percentage of that rail-to-ground voltage to appear rail-to-rail. The following is an example of a field condition where a degraded insulated joint contributed to an interference problem, and the approach used to identify the factors contributing to the problem.
Shorted IJ Example

A railroad experienced apparent interference to the signaling system along a portion of the region of parallel with a power transmission line. The noted interference was restricted to a track segment of approximately six miles, which lies between two power substations, although the exposure extends another several miles south of the second substation. Other substations and taps exist along the exposure. The transmission line is a double-circuit 120-kV vertical arrangement, mostly on lattice-type towers with maximum currents recorded of approximately 900 amperes per phase. For normal operation of the transmission line there are times when interference affects the operation of the train-control and highway grade-crossing track-signal systems. The incidents of signal-system disruption have been observed to occur during periods of hot weather, ostensibly when the load on the power system might be expected to be higher than normal. Two classes of track signaling equipment exist within the region of interference. Coded electronic track circuits are used for block signaling train control. The equipment sends dc pulses through the rails. Within the region between the two substations, there are four track signal locations from south to north at approximately railroad mileposts 18, 19.76, 21.6, and 23.9. The length of the track circuits ranges from approximately 9000 to 12,000 feet. South of MP 18, the track insulators are spaced by shorter distances. One signaled highway/railroad grade crossing exists in the region of apparent interference. The crossing is protected by constant warning time equipment. Track impedance-sensing equipment for grade crossings is typically less tolerant than track-circuit signal equipment to induced power-line voltage. The susceptibility of motion detection and constant-warning time grade crossing equipment is a function of several parameters, including:

The signal function performed i.e., motion sensing, or constant warning time. The manufacturers model and age of the equipment, with the newer models tending to be less susceptible. The operating frequency of the equipment, with the lower frequency equipment, used for higher-speed train operation, tending to be more susceptible.

12-46

Computer Modeling of Joint Corridors

Field Measurements

Railroad personnel had noted interference conditions on the westbound track, but not on the eastbound track. The 60-Hz voltage on the westbound track was measured at three signal locations within the region that had recently experienced signal-equipment problems. Prior voltage measurements on both track in this region showed significantly higher rail-to-rail voltage on the westbound track. Figure 12-40 shows measured voltages at the three signal locations on the westbound track. Figure 12-41 is a sketch showing the location and labeling of the voltage measurements relative to the rails.

Figure 12-40 Measured Voltages at the Three Signal Locations on the Westbound Track

12-47

Computer Modeling of Joint Corridors

Figure 12-41 Sketch Showing the Location and Labeling of the Voltage Measurements Relative to the Rails

At each signal location, Figure 12-40 shows the rail-to-rail voltage north and south of the insulated joint as the two left bars in each group. The voltage across each insulated joint was measured for all except the signal at MP 23.9. At that location, the rail-to-rail voltage north of the joint was not measured, because it was previously found to be a low value. The insulated joint voltage-bars are on the right side of each group in Figure 12-40. The rail-to-rail voltage south of the MP 21.6 IJ was 16 volts, which is in the range that may interfere with the operation of an electronic track circuit unit. The other rail-to-rail voltages for the track circuits north and south of the MP 21.6 are approximately 7 volts. Since significantly higher rail-to-rail voltage was observed on the westbound track than on the eastbound track, a reason for track unbalance was sought. At the MP 21.6 location, significant difference in voltage was measured across the two insulated joints. At the other signal sites shown in Figure 12-40, the voltages across the two IJs were more similar. Measurement of the insulated joint voltage on the eastbound track at the MP 21.6 location showed approximately 40 volts across each insulated joint.

12-48

Computer Modeling of Joint Corridors

The above voltage measurements suggested that the resistance of the two insulated joints of the westbound track at the MP 21.6 location may be considerably different. At a later date the resistances of the westbound track insulated joints at the MP 21.6 and MP 19.76 locations were measured by railroad personnel, using a commercial insulated-joint tester. The results of those measurements are presented in Table 12-1.
Table 12-1 Westbound Track IJ Resistance Measurements (June 2000) Location MP 19.76 MP 21.6 East Rail 7 ohm low West Rail 7 ohm 3 ohm

The east-rail insulated-joint resistance indicated a low resistance range, but a specific value could not be ascertained, because some of the digits were blinking on the instrument. The study used IJ resistances in the above range for most analysis, and a low value (0.2 ohm) for the east rail IJ resistance at the MP21.6. In September 2000, a railroad technician noticed that the east rail IJ at the MP 21.6 location felt warm to the touch, and had the joint replaced. Since the resistance measured by the commercial tester at the MP 21.6 insulated joint was questionable, and may have been caused by the 60-Hz voltage that was present, an alternative measurement procedure was used to evaluate the IJ resistances. A flexible current probe that can be placed around a rail was used by railroad personnel to measure the induced 60-Hz rail current. Figure 12-42 is a photograph of the flexible current probe being used to measure rail current. If the current probe is placed around the rail at an insulated joint location, an accurate measure of the current that flows through the insulated joint can be made. Measurement of the voltage across the insulated joint will permit the magnitude of the insulated joint impedance to be determined by forming the ratio of the measured voltage divided by the measured current. This measurement approach is most easily implemented if the rails are excited by the fields of a power transmission line. When power-line induced voltage is present, no separate source of voltage to force current through the rails is needed. The current and voltage measurements are most accurate if made using a spectrum analyzer, which permits separation of different frequency components of the voltage and current. A broad-band general purpose voltmeter can be used if the 60-Hz component of induced voltage dominates other signal components that may be present.

12-49

Computer Modeling of Joint Corridors

Figure 12-42 Flexible Current Probe Being used to Measure Rail Current

The current-probe measurement approach should result in more accurate values of insulated joint impedance (resistance) than can be obtained with most commercial insulated joint test instruments, because the current-probe approach is not influenced be rail-to-rail connected signal-equipment impedances or by current that flows through the ballast. Railroad personnel used the current-probe approach at the same insulated joints locations shown in Table 12-1, which had been previously measured with a commercially available IJ test instrument. Measurements were made after the MP 21.6 insulated joint was replaced. The current probe indicated less than 0.1 ampere through the joints, for an induced voltage in the range of 30 volts across the insulated joints, which corresponds to resistance well over 100 ohms. The insulated joint that had been cut out was still at the MP 21.6 site during those measurements. The cut-out joint was temporarily jumpered across the new joint and voltage and current measurements were made. The 60-Hz current through the old joint was approximately 9.5 amperes and the voltage was 8.5 volts, corresponding to 0.9-ohm resistance. The two procedures outlined above for measuring rail insulated-joint resistance result in a different range of resistance for the insulated joints. However, these measurements and the voltage measurements of Figure 12-40 indicate an unbalance of insulated-joint resistance for the westbound track at the MP 21.6 location. Power line coupling that results in rail-to-rail voltage at signal-equipment locations can cause an interference condition. The rail-to-rail induced voltage is typically quite small for normal track conditions because of the inherent balance of the track system. The magnetic field from a power-line normally induces a rail-to-earth voltage that is approximately the same on each of the rails. Therefore, some form of unbalance is generally necessary to cause the rail-to-earth voltage to be converted to a rail-to-rail voltage. 12-50

Computer Modeling of Joint Corridors

One of the most significant forms of track unbalance is a degraded or shorted track insulated joint. The normal track signal system is generally insensitive to a single insulated joint with degraded resistance. That is, the track signal system will generally operate normally with a single degraded insulated joint. However, a single low resistance insulated joint can result in conversion of power-line induced rail-to-earth voltage into rail-to-rail voltage, which can interfere with the signal-system operation.
Analysis of Exposure

The purpose of the analysis was to identify factors that can contribute to the incidence of interference to the railroad signal systems, and to identify remedial measures that can reduce or eliminate the incidence of interference. The approach used a computer model of the power and railroad systems to evaluate the effect of important variables that may influence interference to the railroad signal systems. The initial portion of the study included a review of the power and rail-system information and measurements on the rail system, and the development of a suitable model for analysis of electromagnetic interaction along the corridor. The influence of steady state induction on signal equipment performance is dependent on the specific equipment; the arrangement of the track circuits; and operating conditions that may result in an unbalance of the track system. For this investigation, limited steady state 60-Hz rail voltage measurements and rail insulated joint resistance measurements were made and used to compare to the preliminary model calculations. Those baseline analyses were used to adjust other unknown parameters to obtain reasonable agreement between the model results and measured values. The model was then used to evaluate the effect of selected mitigation measures on the induced voltage.
Model Comparison to Field Measurements

The two mainline rails of the rail system were modeled from one block north of the MP 23.9 signal location south to MP 12.8. Within this region, all track insulators were modeled. The railto-rail impedance of the coded track circuit equipment was modeled at all signal locations from MP 17.2 to MP 24.16. The power line was modeled at a point 50 feet west of the eastbound track from MP 13 to MP 21.6, where it crosses to the east of the track. The power line was modeled at 50 feet east of the westbound track between MP 21.6 and MP 22.6. Between MP 22.6 and 23.3 the transmission line increases distance from the track in the vicinity of a substation feed. Between MP 23.3 and 23.8 the transmission line is approximately 50 feet east of the westbound track. North of that location, the transmission line crosses to the west of the tracks on route to a substation, which is on the west side of the tracks. As noted, there are several power substations along the common corridor. The current in the transmission line and the phase relationship between transmission line conductors is different in the segment between each substation. 12-51

Computer Modeling of Joint Corridors

The transmission-line currents were modeled as 5 segments between the various substations, based on an estimate of the peak current during June 2000, when the track-voltage measurements were made. It is unlikely that the transmission line current was at the recorded maximum. We do not know the transmission line currents during the period when track voltage measurements were made. The transmission-line currents were likely less than the values used for the analysis. In addition, we have no measure of the track ballast or soil resistivity when the rail voltages were measured. We have assumed 100-ohmm soil, which was suggested by the power company as representative for the area. We used a track ballast resistivity of 10 ohmkft as a representative value. Calculation was made with the model to compare to the measured track voltages. The track insulated joints were modeled as 7 ohms, except at the MP21.6 location. At that location, the east rail IJ of the westbound track was modeled as 0.2 ohms, and the other insulated joint was modeled as 3 ohms, consistent with the IJ resistance measurement of Table 12-1. At the time of the calculations, the defective cut-out IJ resistance measurement with the flexible current probe was not available. However, the calculated results are not expected to change significantly for shorted IJ resistance between 0.2 and 0.9 ohms. Figure 12-43 shows the calculated rail-to-rail and insulator voltages for the westbound track with a shorted insulator (0.2 ohms) at the MP 21.6 location. The model results of Figure 12-43 shows that the calculated rail-to-rail voltages, particularly at the MP 21.6 signal where problems are observed, agree well with the measured values of Figure 12-40. Thus, the basic model appears to be in reasonable agreement with the physical and electrical arrangement within the exposure. The model appears sufficiently realistic to use to assess the effects of mitigation measures on the predicted rail-to-rail voltage. There is a signaled grade crossing at approximately MP 22.18. The calculated westbound track rail-rail voltage at MP 22.18, is approximately 7 volts for the modeled degraded insulated joint at the MP 21.6 signal-location, as in Figure 12-43. This voltage level should interfere with most grade crossing equipment and railroad personnel had noted incidents of malfunction of the crossing equipment at that location. Figure 12-44 shows the calculated rail voltages for the same conditions as Figure 12-43, but without a degraded insulator at the MP 21.6 signal location. That is, with balanced resistance at all pairs of insulated joints. Figure 12-44 shows that the balanced IJ resistances results in significantly lower calculated rail-to-rail voltages, than for the degraded insulator condition of Figure 12-43. Voltage measurements made by railroad personnel at the MP 21.6 signal location, after replacing the degraded east-rail IJ in the westbound track, compared well with the calculations of Figure 12-44. The measured rail-to-rail voltages were 2-volts north of the IJ and 0.8-volts south of the IJ. The measured voltages across the IJs were both approximately 37 volts, which is very similar to the calculated values shown in Figure 12-44. The balanced IJ conditions of Figure 12-44 resulted in a calculated rail-to-rail voltage of 1.5 volts at the MP 22.18 grade-crossing location, which is a significantly lower value than with unbalanced IJ resistances at the MP 21.6 signal location.

12-52

Computer Modeling of Joint Corridors

Figure 12-43 Calculated Rail-to-Rail Voltages

Figure 12-44 Calculated Rail Voltages for the Same Conditions as Figure 12-43, but Without a Degraded Insulator

12-53

Computer Modeling of Joint Corridors

The foregoing analysis has illustrated that a degraded rail insulator can unbalance the track circuit resulting in rail-to-rail induced voltage. The analysis shows that 60-Hz magnetic-field excitation by the existing transmission line configuration results in rail-to-rail voltages that are:

consistent with observed values. in the range to cause interference to the existing signal system.

Power-System Phasing

The phasing of the double-circuit transmission line between the substations at MP 17 and MP 24.5 results in the highest field excitation for a parallel railroad facility of all possible phasing arrangements. However, the phasing arrangement that is used south of the MP 17 substation results in the lowest fields of any possible phasing arrangement. The effect of changing the phasing of the segment of north of MP 17 to be the same as is presently used south of that location was investigated by analysis. Figure 12-45 shows the calculated rail voltages at signal locations for conditions of:

Reversed transmission-line phasing for the MP 17-to MP 24.5 region (xyz-zyx top-tobottom), and estimated dissimilar circuit currents. A degraded rail insulator at the MP 21.6 signal that resulted in Figure 12-43 for the existing transmission line phasing.

Figure 12-45 Calculated Rail Voltages

12-54

Computer Modeling of Joint Corridors

The display of Figure 12-45 shows the rail-to-rail voltages as the two left-side bars at each signal location. It is seen that the predicted rail-to-rail voltage is less than 5 volts at all signal locations. Figure 12-45 shows that modifying the phasing of the MP 17-to-24.5 segment of the transmission line results in predicted rail-to-rail voltage for a shorted rail insulated joint at the MP 21.6 signal that is significantly less than for the present phase arrangement. With the modification in the phasing of the transmission line, the signal-system should be compatible, even if a degraded IJ occurs in the future. The voltage that will appear rail-to-rail at the input terminals of signal-equipment as a result of a degraded insulated joint will depend on at least the following parameters:

The magnitude of the power-line field that excites the rails. The lengths of the track circuits on each side of the degraded joint. The resistance of the degraded joint. The track ballast resistivity. The soil resistivity. The location of the signal equipment in the track circuit. The impedance of the signal equipment at the frequency of the induced voltage.

From a practical standpoint, the purpose of modeling a specific exposure is not to accurately model the value of each of the above parameters. Generally too many of the variables are unknown. The purpose of the analysis is typically to determine the range of induced voltage that may occur, if the variables tend to be at or near their worst-case values. Therefore, an analysis tries to identify reasonable values for the different variables that tend to make the composite result near a maximum. The important characteristic of the degraded rail insulated joint is that it makes the effective length of one rail longer than the other. That occurrence causes the rail-to-ground voltage, as a function of location, to be different along the two rails. Since the rail-to-rail voltage is the difference between the rail-to-remote-earth voltage at each position along the rails, inequality of the rail-to-ground voltage results in rail-to-rail voltage. Other conditions can also result in inequality of the voltage on the two rails. Since the two insulated joints of a pair can be separated by several feet, a slow-moving train that traverses the insulated joint stagger region can causes one rail to appear electrically much longer than the other during the time that it takes the train to cover the stagger distance, also resulting in unequal induced voltage on the two rails, and thus a 60-Hz or harmonic voltage rail-to-rail. The induced rail-to-rail voltage caused by a train moving through the insulated-joint stagger can be similar in magnitude to the induced voltage caused by shorted IJ, and can influence the operation of grade crossing signal equipment. Degraded IJ with Train The previous paragraphs have described how a degraded insulated joint can result in unbalance of the track circuit, so that power-line induced voltage can appear at the input to signal equipment that is connected rail-to-rail. One purpose of evaluating the voltage that can result from a degraded insulated joint is to evaluate methods that may be available to help reduce 12-55

Computer Modeling of Joint Corridors

the voltage or to control the voltage to a tolerable value. Since unbalance of the track system by a degraded insulated joint is an event that may be difficult to prevent, evaluation of approaches to minimize the effect of a degraded IJ can be performed as part of a compatibility investigation.
Field-Measured Example

Field investigations have shown that the presence of a train in one of the track circuits can increase the rail-to-rail induced voltage that occurs as a result of a degraded rail insulated joint. Figure 12-46 shows several sketches. The top sketch indicates two track circuits, with three pairs of insulated joints that are identified at locations A, B, and C. A unidirectional grade-crossing predictor (constant warning time unit) existed on each side of the IJ at location B. The investigation was made to evaluate the effect of a degraded (shorted) insulated joint at location B on the rail-to-rail voltage induced by a power transmission line. An oscillographic pen recorder was connected rail-to-rail on each side of the insulated joint at location B to monitor the rail-to-rail voltage as a passenger train moved through the track circuits in the direction of from A to B to C, from west-to-east. The additional sketches in Figure 12-46 show the location of the train at successive increments of time, identified as Time 1 (just crossing the IJ at location A) to Time 5 (just crossing the IJ at location C).

Figure 12-46 Track Sketches

12-56

Computer Modeling of Joint Corridors

Figure 12-47 is a sketch of the resulting oscillograph recording of the rail-to-rail voltage on each side of the insulated joint at location B as the train moved through the region. The rail-to-rail voltage on the west side of the IJ at location B is denoted Vw, and the voltage Ve is the rail-to-rail voltage east of the IJ. Figure 12-47 also identifies the same five times that the train location was shown in Figure 12-46.

Figure 12-47 Oscillograph Recording of the Rail-to-Rail Voltage on Each Side of the Insulated Joint

Before Time 1

The train is outside of the two-block region defined by insulated joints A, B, and C. The rail-to-rail voltages Vw and Ve are at unoccupied track values caused by the power line field excitation and the presence of a shorted insulated joint at location B.
At Time 1 The front of the train has crossed over Insulated Joint A.

The voltage Vw suddenly decreases as the front of the train crosses over the IJ at Location A, because of the low rail-to-rail impedance which forces the two rails to the same potential on the east side of the IJ at Location A. As the train moves farther east, the potential difference between the rails becomes less.

12-57

Computer Modeling of Joint Corridors

The voltage Ve suddenly increases as the front of the train crosses over the IJ at Location A, because of the low rail-to-rail impedance at Location A which increases the voltage on the shorted rail. The highest value of Ve occurs for the train to the west of the IJ at Location B.
At Time 2 The front of the train has crossed over Insulated Joint B.

The voltage Vw is zero because the two rail are shorted together by the train on the west side of the IJ at Location B. The voltage Ve suddenly goes to zero as the train shorts the two rails together on the east side of the IJ at Location B.
At Time 3 The end of the train has crossed over Insulated Joint B.

The voltage Vw suddenly increases because no part of the train is in the track circuit between IJ A and IJ B. The voltage Ve is zero, but starts to increase as the end of the train moves farther east of the IJ at Location B.
At Time 4 The front of the train has crossed over Insulated Joint C.

The voltage Vw suddenly increases because the effective length of the shorted rail (and the location of the zero voltage point) increases because the train bridging across the IJ. The voltage Ve continues to increase as the train shunt moves farther from the IJ at Location B.
At and after Time 5 The end of the train has crossed over Insulated Joint C.

The voltage Vw suddenly decreases because the effective length of the shorted rail decreases because the train no longer bridges across the IJ at Location C. The voltage Ve suddenly increases as the end of the train crosses over the IJ at Location C, because the train is not providing a shunt in the track circuit between IJs at Locations B and C.
This example has shown that the highest rail-to-rail voltage with a degraded IJ is not with the unoccupied condition. The highest rail-to-rail voltage on one side or the other of a degraded insulated joint occurs when a train provides a short across an IJ that is adjacent the degraded IJ. If analysis is to be performed to assess the worst-case voltage that will be impressed at signalequipment locations for normal operation of the railroad, the analysis of a degraded IJ with a train-shunt at an adjacent IJ should be included in the analysis.
Simple Model with Shorted Insulator

The simple geometry of Figure 12-46 has been analyzed for a constant distance of 5000 feet between insulating joints to illustrate the predicted effect of a train shunt along with a degraded track insulated joint. The model includes track circuits that are west and east of the region that is shown in Figure 12-46. Negligible impedance loading of the track by signal equipment impedance is assumed for the analysis. A representative transmission line geometry was assumed for the analysis. The results have been normalized with respect to the longitudinal electric field at 12-58

Computer Modeling of Joint Corridors

the track location, which tends to make the results independent of the details of the transmission line. The induced voltage is symmetrical with respect to which side of the shorted insulator the train shunt is located for equal-length track circuits. Four conditions have been analyzed, which only include a train shunt on one side of the shorted insulator location because of the symmetry. Figure 12-48 shows the calculated normalized rail-to-rail voltage at the IJ locations A, B, C, and D (which is the next IJ east of location C) of the Figure 12-46 geometry for the following four conditions:

Normal unoccupied track circuits, with no degraded insulating joints. (this condition is the left-most bar at each IJ location in Figure 12-48). A shorted rail insulator at location B without train. (this condition is the second from the left-most bar at each IJ location in Figure 12-48) A time just prior to time 4 in Figure 12-46, for which the train has not crossed the insulating joints at location C. (this condition is the third from the left-most bar at each IJ location in Figure 12-48) The time 4, after the front of the train crosses the insulating joints at location C. (this condition is the right-most bar at each IJ location in Figure 12-46)

Rail-to-Rail Voltage [V/(V/km)] 2

Ballast = 100 Ohm-kft Distance Between IJ's = 5000 ft 1.5 Unoccupied Track & Normal IJ's Unoccupied Track & Shorted IJ @ B Occuped West of IJ C & Shorted IJ @ B Occupied Across IJ C & Shorted IJ @ B

0.5

0 Ae Bw Be Cw Ce Dw

Figure 12-48 Calculated Induced 60 Hz Rail-to-Rail Voltage at Insulated Joint Locations for Selected Operational Conditions

12-59

Computer Modeling of Joint Corridors

The calculated rail-to-rail voltages are shown for an assumed 100 ohmkft ballast resistivity and 100 ohmm soil resistivity. The rail-to-rail voltage in Figure 12-46 has been normalized by the longitudinal electric field (LEF) at the track location. The key results for the four analyzed conditions are:

The normal track condition results in very low rail-to-rail voltage at all locations. The shorted insulator condition results in rail-to-rail voltage that is approximately the same at locations A, B, and C. A train shunt on the west side of location C results in higher rail-to-rail voltage at locations on the west side of the shorted insulator, and lower voltages on the east side because of the train shunt. This is consistent with the recorded voltage in Figure 12-47, which shows the voltage Vw is somewhat higher with the train at Location 3 than with no train. A train shunt that spans the track insulators at location C, results in a further increase in the voltage in the region bounded by the insulators at location A and B. This is consistent with the recorded voltage in Figure 12-46, which shows the voltage Vw is the highest value between times 4 and 5, for which the train shorts across the IJs at Location C.

Figure 12-49 shows the results for the same four analyzed conditions, except with the ballast resistivity reduced to 3 ohmkft. The resulting rail-to-rail voltage from a shorted insulator is less than for the higher ballast resistivity of Figure 12-48. In addition, the voltage increase that is caused by a train shunt is not as significant as for the higher ballast resistivity.
Rail-to-Rail Voltage [V/(V/km)] 2

Ballast = 3 Ohm-kft Distance Between IJ's = 5000 ft

1.5

Unoccupied Track and Normal IJ's

Unoccupied Track & Shorted IJ at B

Occupied West of IJ C & Shorted IJ @ B


1

Occupied Across IJ C & Shorted IJ @ B

0.5

0 Ae Bw Be Cw Ce

Figure 12-49 Calculated Results for the Same Four Analyzed Conditions of Figure 12-48, Except with the Ballast Resistivity Reduced to 3 OhmKft

12-60

Computer Modeling of Joint Corridors

The rail-to-rail voltages shown in Figure 12-48 and Figure 12-49 are at the insulated joint locations. Track occupancy signal equipment will be located at insulated joint locations. Highway grade crossing motion detection and prediction equipment may also be located at insulating joints; however, grade crossings may be at locations where insulated joints do not exist. Therefore, for that type of equipment, the rail-to-rail voltage at the actual locations where the grade crossings exist is of importance for compatibility assessment. Signal Equipment Impedance Track signal equipment is generally connected rail-to-rail. Personnel safety issues are typically not influenced by signal equipment that is connected rail-to-rail. However, some equipment such as Wide-Band Shunts and Tuned Joint Couplers are connected across insulated joints and provide an impedance that couples between two track circuits. The impedances that appear across an IJ can influence the rail-to-ground voltage, and thus may influence personnel safety. However, the primary concern with regard to signal equipment impedances is that induced voltage can appear across the impedance and may influence the normal operation or performance of the signal equipment. That is, the induced voltage may interference with the normal operation of the signal system. The impedance of signal equipment is dependent on the type of equipment. The 60-Hz and harmonic impedances of most railroad signal equipment is not readily available. The manufacturers do not publish impedance values, and may not typically measure the impedance at power frequencies. A common track circuit is the dc battery and relay track circuit. At the power frequency a battery is a very low impedance, however, a dc track circuit usually has a resistance in series with the battery for limiting the current output and to adjust the current through the relay. This resistance is typically an ohm or so. A choke may also be in series with the battery to prevent gradecrossing frequency signals from being affected by the battery. Such a choke is also a high impedance for power frequencies. DC track circuit relays typically range from -ohm to 4 ohms. However, at 60 Hz these relays have a high inductive impedance, typically 100 ohms or more. The impedance magnitudes of many electronic track-circuit units and grade crossing motion detectors and constant warning time devices are a few ohms. The phase angle can range from nearly +90 degrees to nearly -90 degrees. Narrow band shunts (NBS) and tuneable joint couplers (TJC) are basically series-tuned inductance and capacitance (LC) circuits that are tuned to the frequency associated with a highway grade crossing, such as a motion detector or a constant-warning time device. Since the grade crossing units operate at higher frequencies than the power line frequency, the tuned frequency of NBS or TJC circuits are also higher than 60 Hz. Therefore, since an LC circuit is capacitive below the tuned frequency these tuned circuits are capacitive at the power-line frequency. However, at harmonic frequencies the tuned-circuit impedance may be either inductive or capacitive. The magnitude of the TJC or NBS 60-Hz impedance ranges from about 10 ohms for the lower grade-crossing frequencies, to up to approximately 200 ohms for high grade-crossing device frequencies.

12-61

Computer Modeling of Joint Corridors

The signal-equipment impedances that are connected rail-to-rail will influence the induced railto-rail voltage that occurs as a result of a particular unbalancing mechanism. As an example, we have calculated the rail-to-rail induced voltage for a simple two track-circuit arrangement, with a degraded insulated joint, as a function of the impedance magnitude of signal equipment that is connected rail-to-rail at the ends of the track circuits. The results of the example are shown in Figure 12-50. The figure shows a sketch of the track circuits under consideration. The sketch shows one of the insulated joints as a resistance. The analysis for this example has assumed a 0.1-ohm resistance insulated joint. Impedances are also shown at each end of each track circuit to represent the impedance of signal equipment. There are two curves in the figure. One curve (V1) is the rail-to-rail voltage at the location of the degraded insulated joint. The other curve (V2) is the rail-to-rail voltage at the other end of the track circuit. The voltage curves are normalized by the rail-to-ground voltage induced with nonperturbed (all good) insulated joints, and expressed as a percent. The voltage curves are plotted as a function of the signal equipment impedance magnitude. For the analysis, the equipment impedance was assumed to be an inductive reactance, which has an impedance angle of 90 degrees. The curves show that the lower the impedance of the signal equipment, the lower the induced rail-to-rail voltage. For a representative value of 10 ohms for the equipment, the rail-torail voltage at the shorted IJ location is approximately 65% of the voltage that would appear railto-ground at that location, if the insulated joint was not shorted. Thus, for this example, if the rail-to-ground voltage is 10 volts with good IJs, the voltage that would appear across the equipment with a shorted insulated joint is expected to be approximately 6.5 volts. That value of rail-to-rail induced voltage will not adversely affect some electronic track circuits, but may interfere with some highway grade-crossing units, for which AREMA guidelines recommend a 5-volt interference tolerance [80].
100 90

Rail-to-Rail Voltage as Percentage of Nonperturbed Rail-to-Ground Voltage (%)

80 70 60 50 40 30

V1, Rail-Rail Voltage at Degraded IJ 1.5 mile track circuits 100 ohmkft ballast 100 ohmm soil 0.1 ohm degraded IJ 100 ohm good IJ Angle of impedance +90

V2, Rail-Rail Voltage at Adjacent IJ

V1

V2

Degraded IJ

Adjacent Normal IJ

20 10 0 1 10 100

Magnitude of Equipment Impedance (ohms)

Figure 12-50 Affect of Equipment Impedance on Induced Rail-to-Rail Voltage

12-62

Computer Modeling of Joint Corridors

Insulated joints that occur within a highway grade-crossing approach must be bypassed to allow the grade-crossing device signal to flow past the insulated joint so the signal device can see all the way to the end of the approach. Tuneable joint couplers and wide-band shunts are two devices that are used to allow grade-crossing systems to see past the insulated joint. Tuneable joint couplers are placed across each insulated joint of a pair. They will tend to have the same effect on the induced rail voltage as a pair of degraded insulated joints. Figure 12-34 and 12-35 showed the magnitude of the calculated rail-to-ground voltage that resulted from a 30 or 10 ohm insulated joints at one location. Since some tuneable joint couplers are in this impedance range for 60-Hz induced voltage, similar patterns on the rail voltage might be anticipated for tuned-joint couplers at one IJ site. The result is that the rail-to-ground voltage and the voltage across the insulated joint are reduced at the site of the TJCs but are increased at adjacent insulated joint locations. The power line induced voltage that is developed across tuned joint couplers is of concern, since the magnetic cores that are used for the inductors can change permeability, and thus change inductance, if the induced power line-voltage becomes too high. For some tuneable joint couplers, the 60-Hz or 180-Hz voltage that will start to detune the TJC can be in the low-volts range (less than 10 volts), depending on the type, manufacturer, and tuned frequency. Since the induced voltage that is developed across insulated joints is in the range of twice the rail-toground voltage, almost any significant power line exposure may present a problem for grade crossings that have tuned-joint couplers within the exposure. Wide-band shunts are very high value capacitors that are used to bypass rail insulated joins so that the grade crossing signal(s) can pass from one track circuit to another. Wide-band shunts (WBS) are also sometimes used as a grade-crossing approach termination, for which they are connected rail-to-rail. The capacitance of wide-band shunts is typically in the range of 100- to 300-thousand microfarads. That is, between 0.1 and 0.3 farads. The impedance of these devices at 60 Hz is in the range of 10 milliohms, which is quite a small impedance. A wide-band shunt across an insulated joint is approximately the same as a continuous rail. Therefore, two track circuits that are separated by insulated joints that are bypassed by WBSs tend to have induced voltage similar to a single effective track circuit with a length that is the sum of the two circuit lengths. Therefore, tuned joint couplers or WBS across IJs are typically a problem for powerline exposures, and often the signal system needs to be re-arranged to prevent interference problems. Shorted Arrester to Ground Track circuit lightning arresters are designed to fail as an open circuit. However, sometimes they fail as a low-impedance, which results in one rail being connected to the grounding system (made ground) of the signal-equipment enclosure. The arrangement of the made ground for the signal-equipment enclosure varies with the type of enclosure and the individual railroad grounding practices. Often the made ground consists of four ground rods, one at each corner of the enclosure, that are connected together by buried bare copper conductors (often #6 AWG). The goal is to obtain a resistance of the made ground that is 15 ohms or less [81]. If one of the rail-to-ground lightning arresters becomes shorted, one rail is connected to the signal-equipment enclosure made ground at that location. The connection of rail-to-ground 12-63

Computer Modeling of Joint Corridors

may unbalance the track impedance so that power-line induced voltage is not the same on both rails, thus causing a rail-to-rail voltage at the power frequency or harmonics. Interference may result if the induced rail-to-rail voltage is sufficiently high at the location of sensitive railroad signal equipment. A shorted arrester to ground generally only affects the induced voltages in the same track circuit as the shorted arrester. The induced voltage that will be developed between the two rails depends on several factors, including:

The magnitude and frequency of the exciting longitudinal electric field from the power line, The length of the track circuit in witch the arrester to ground fails, The impedances of all the signal equipment in the track circuit at the frequency of the induced voltage, The resistivity of the soil and track ballast, The resistance to ground of the signal equipment prime ground connection.

These factors must be adequately represented in the model to evaluate the induced voltage at the input to interference-sensitive signal equipment. Of course, for most exposures to be assessed by analysis, such detailed physical and electrical values are not always available. Therefore, estimates for typical values, or expected extremes in important parameters are often used in an attempt to bound the results to a range of values that could occur.
Affect of Signal-System Ground Resistance on Coupled Interference

A shorted track arrester connects the rail to the signal equipment enclosure prime ground, which is connected to the made ground (grounding electrode) of the signal enclosure. However, the prime ground is often also connected to the neutral of the local distribution line that supplies power to the signal-equipment within the enclosure. Therefore, the resistance to ground of the signal equipment enclosure prime ground is the parallel impedance of

the grounding electrode of the signal equipment enclosure the distribution line neutral.

The total resistance of the prime ground may be lower than the resistance of the made ground of the enclosure. The neutral of a modern distribution line is a multi-grounded neutral, which is typically connected to ground at least at four locations per mile, and is also grounded at each service drop, where connections may be made to other low-resistance underground objects such as metal underground water pipes, metal frames or buildings and other types of grounding electrodes permitted by the National Electric Code. The resistance looking into the distribution line multi-grounded neutral can be a low value of resistance, which can affect the equivalent resistance of the signal-equipment enclosure prime ground and the induced rail-to-rail voltage. Figure 12-51 shows how the calculated rail-to-rail voltage:

at a shorted arrester to ground (V1), and at the other end of the track circuit (V2)

12-64

Computer Modeling of Joint Corridors

vary as a function of the resistance to ground. The example assumes that the shorted arrester is associated with the signal equipment at the ends of a track circuit. A sketch within the figure shows the track circuits under consideration. For the analysis, we assumed that the track circuits were uniformly excited by the 60-Hz longitudinal electric field of a power transmission line. In the absence of the shorted arrester, the power line field excitation would cause a rail-to-ground voltage. The rail-to-rail voltage curves in Figure 12-51 have been normalized by this nonperturbed rail-to-ground voltage (expressed as a percentage) so that the results are independent of the exciting power line field.
60

50

V1 Value at the site of the shorted arrester

Rail-to-Rail Voltage as a Percent of Nonperturbed Rail-to-Ground Voltage

40

30

1.5 mile track circuits 100 ohmkft ballast 100 ohmm soil 100 ohm IJ's Equipment impedance j10 ohms

V1

V2

20
V2 Value at the distant end of the same track i it

10

0 0.1 1 10 100

Shorted Arrester Resistance

Figure 12-51 Rail-to-Rail Voltage with a Shorted Track Arrester, Normalized by Non-Perturbed Rail-to-Ground Voltage

The curves of Figure 12-51 show that the induced rail-to-rail voltage becomes a higher percentage of the non-perturbed rail-to-ground voltage (the rail-to-ground voltage without a failed arrester) for lower values of grounding resistance. That is, as one rail becomes better grounded the difference between that rail and the other rail voltage becomes greater. For the chosen parameters of this example, if the grounding resistance is just a 15-ohm made-ground resistance, the rail-to-rail voltage at the shorted arrester location is approximately 15% of the rail-to-ground voltage that would exist without the shorted arrester. However, if the grounding resistance is as low as 2 ohms, perhaps because of the effect of the parallel resistance of the power distribution multi-grounded neutral, the rail-to-rail voltage at the shorted arrester location is over 40% of the rail-to-ground voltage that would exist without the shorted arrester. Figure 12-52 illustrates how the rail-to-rail voltage varies over the length of the track circuit for the parameters of this example, for two different values of grounding resistance. The rail-to-rail voltage gradually decreases with distance away from the shorted arrester location. 12-65

Computer Modeling of Joint Corridors


45
Shorted Arrester to 2 ohm Ground

40
Shorted Arrester Location End of Track Circuit Opposite End of Track Circuit

Rail-to-Rail Voltage Percent of Nonperturbed Rail-to-Ground Voltage

35

30

25

1.5 mile track circuits 100 ohmkft ballast 100 ohmm soil 100 ohm IJ's Equipment impedance j10 ohms
Shorted Arrester to 15 ohm

20

15

10

Track Circuit

0 -3 -2 -1 0 1 2 3 4

Distance from Shorted Arrester in Miles

Figure 12-52 Calculated Rail-to-Rail Voltage Profile Over the Length of the Track Circuit

The above example has considered the ballast resistance to be a rather high value (100 ohmkft). The induced voltage is less for lower values of ballast resistance, which might be more normal. As an example to compare to the 2-ohm ground case described above, if the ballast resistance is 10 ohms, then the rail-to-rail voltage at the shorted arrester location is approximately 26% of the rail-to-ground voltage that would exist without the shorted arrester. Thus, lower ballast resistance tends to result in less chance of interference for the same power line exposure with a shorted track arrester for two reasons:

the induced rail-to-ground voltage without a shorted arrester is lower for the lower ballast resistance, the percentage of the rail-to-ground voltage that is converted to rail-to-ground voltage is also lower for the lower ballast resistance.

However, ballast resistance can change significantly with weather conditions. Over time a track circuit may experience a wide range of ballast resistivity. The above results suggest that since high ballast is the worst case for shorted-arrester interference, the high-ballast conditions should be considered, rather than the more normal ballast conditions. Determining an appropriate value of ballast resistivity to use for the high-ballast condition is less obvious. Generally, increases in ballast resistivity have little effect on induced voltage for ballast resistivity that is higher than approximately 100 ohmkft. Since 100-ohmkft ballast can occur for hot-dry conditions or frozen conditions, at least for good ballast, it seems reasonable to consider that range of ballast resistivity for predicting the rail-to-rail voltage that might be expected to result from a shorted rail-to-ground arrester. 12-66

Computer Modeling of Joint Corridors

Conduction Coupling to Track Circuit

The foregoing discussion of shorted arrester interference has focused on track circuits that are excited by the inductive coupling from power lines. However, a shorted rail-to-ground arrester can also help cause interference to be coupled into the rail system with no inductive-coupling exposure. The discussion above has noted that the ground of the signal-equipment enclosure is often common to the power distribution-system multi-grounded neutral. The distribution system neutral is not necessarily at zero voltage. The distribution system neutral can be at voltage for several reasons including:

IR drop in the neutral because of high return current from single-phase loads, Inductive coupling between the distribution-line neutral and phase conductors may cause peaks in voltage on the neutral near locations where loads are dropped, that is at locations where the exiting field at the neutral changes significantly, Inductive coupling between the distribution-line neutral and phase conductors of an overbuilt transmission line may cause peaks in voltage on the neutral near locations where the coupling changes (e.g. the distribution line moves off the transmission line structures). Breaks in the neutral, which can result in high neutral voltage and voltage at ground connections to the neutral.

The voltage that can occur on a distribution system with a broken multi-grounded neutral is one of the causes of the stray-voltage that is of concern to dairy farmers. To help quantify this effect, Figure 12-53 shows a sketch of rail track circuits and one rail-to-ground shorted arrester that is connected to the signal-system ground Rg. The signal system ground is also assumed to connect to the distribution line neutral, which for this example appears as a resistance Rd to ground at the connection point. Voltage that is induced onto the neutral by the power system can be represented as a current source, which is shown in, Figure 12-53. Assume for example that the equivalent distribution system current source causes a voltage Vg at the grounds (the parallel combination of signal enclosure and distribution system grounds) when the arrester is not shorted. The rail-to-rail voltage that will occur if an arrester shorts to ground is dependent on the track-circuit length and the ballast resistivity.

12-67

Computer Modeling of Joint Corridors

Figure 12-53 Sketch of Track Circuit for Coupling of Interference from Distribution Neutral Through Shorted Track Arrester

Table 12-2 provides representative calculated values of the rail-to-rail voltage caused by conducted coupling between a distribution line and a track with a shorted arrester to ground for 1.5-mile track circuits. The rail-to-rail voltage is expressed as a percentage of the voltage on the signal ground without the shorted arrester. The table illustrates that for a given voltage on the distribution neutral, higher rail-to-rail voltage results from higher track-ballast resistivity and lower values of power-system neutral impedance. Of course, lower values of power-system neutral impedance generally results in lower voltage on the neutral.
Table 12-2 Conduction Voltage Coupling Conversion of 60-Hz Voltage on Signal Ground into Railto-Rail Voltage by Shorted Track Arrester (15-ohm Signal Ground Resistance) Percent Conversion of Ground Voltage to Rail-to-Rail Voltage Ballast Resistivity (ohmkft) 10 ohmkft 100 ohmkft Power Neutral Impedance = 1 ohm 25% 40% Power Neutral Impedance = 15 ohm 4.8% 13%

To interfere with grade-crossing signal equipment (for example motion detection or constantwarning time devices) that meet the AREMA-recommended 5-volt interference tolerance would necessitate more than 10 volts from neutral to ground for the high-ballast and low neutralimpedance condition, and considerably more for the low-ballast condition. Of course, if the signal device is susceptible to much lower power-frequency voltage levels, then correspondingly less distribution line neutral to ground voltage may cause interference.

12-68

Computer Modeling of Joint Corridors

Measurement of the voltage on power distribution system neutrals has been made at many farm locations by two groups in Wisconsin organized for the investigation of stray voltage problems. One group is the Wisconsin Stray Voltage Analysis Team (SVAT), the other is the major investor-owned utilities (IOUs). Data collected by these groups has been summarized in a report by the Public Service Commission of Wisconsin (PSCW) [82]. Figure 12-54 is data reported in reference [75], which shows that of the approximately 200 neutral voltages measured on the secondary side of distribution transformers, approximately 2.5% exceeded 4 volts2. Similar data do not exist for the power feeds at railroad signal locations, but if the data of Figure 12-54 is representative at rural railroad sites, the possibility of interference to motion detection or constant-warning time grade-crossing equipment by this coupling mechanism appears to be of low, but finite probability.
60 53 50

46

40 % of Farms

36 32

SVAT IOU

30

20

10

10

9 5 3 3

2 more

0 0-1 1.1-2 2.1-3 Range - Vneutral VOLTS 3.1-4

Figure 12-54 Secondary Neutral to Earth Voltage [75]

Analysis Procedure for Evaluating Induced Voltage at Track Signal Equipment During the development of the computer model of a rail system that will be exposed by a proposed power transmission line, the location of all insulated joints and rail-connected signal equipment is identified, along with the type of equipment. The goal is to include in the model all signal equipment impedances that may affect the induced rail-to-rail voltage, and to identify the locations where rail-to-rail voltages (to assess equipment compatibility with the predicted
2

Elevated neutral-to-earth voltages on farms come from two sources: power line neutrals and on-farm sources.

12-69

Computer Modeling of Joint Corridors

voltages), as well as voltages across IJs, will be calculated. The operating conditions of the power line are reviewed to identify near-worst-case steady-state operating conditions, which will produce the highest fields along the railroad exposure. The normal rail condition generally results in a relatively small rail-to-rail voltage, because of the inherent balance of the tracks, which is similar to a lossy parallel-conductor electrical transmission line. However, unbalances in the track system do occur, and are considered by many railroad personnel to be normal operational conditions. Therefore, a compatibility study should include an evaluation of representative track unbalances to determine if the resulting induced rail-to-rail voltages are compatible with the signal system, and if not, how mitigation alternatives might affect the induced voltage. Typically the unbalancing conditions that are evaluated include:

A degraded or shorted track insulator, accompanied by a train shunt at an adjacent IJ. This is evaluated for each IJ. A degraded or shorted track lightning arrester at each signal location, one at a time.

Normally the procedure used to evaluate the degraded (or shorted) track arrester is to systematically place a shorted insulated joint at each possible insulated joint location. Usually a shorted IJ is only placed at one joint of a pair. For each shorted IJ location, two analyses are performed:

First, the next up-stream IJ pair from the one that is shorted is bypassed (shorted around), along with a 60-milliohm rail-to-rail shunt at the up-stream IJ location. For that condition, the voltage across each signal-equipment impedance, including narrow-band shunts and tuned joint couplers, is calculated and stored. Second, the next down-stream IJ pair from the one that is shorted is bypassed (shorted around), along with a 60-milliohm rail-to-rail shunt at the down-stream IJ location. For that condition, the voltage across each signal-equipment impedance, including narrow-band shunts and tuned joint couplers, is calculated and stored.

At each location where a rail-to-ground lightning arrester exists a shorted IJ to the signalequipment ground is modeled. For that condition, the voltage across each signal-equipment impedance, including narrow-band shunts and tuned joint couplers, is calculated and stored. Nominally, if there are (n) IJs and (m) grade-crossing locations, the number of analysis runs needed to evaluate the track unbalance is (2n) analysis runs for shorted IJs, plus (n+m) analysis runs for shorted arresters, or (3n+m) analysis runs. As noted above, the effect of a track unbalancing condition is characterized by recording the predicted rail-to-rail voltage at each block signal and grade-crossing signal-equipment location plus the voltage developed across each grade-crossing termination shunt and each tuned-joint coupler. Generally, the largest value (maximum) of the voltage at each of those locations is identified across the set of all (3n+m) unbalancing conditions. This process is repeated for each power system condition of concern. The calculated maximum voltage at each signal equipment can be compared to the expected susceptibility threshold for each type of equipment to assess the locations where incompatibility may occur and mitigation measures may be needed.

12-70

Computer Modeling of Joint Corridors

Susceptibility of Track Signal Equipment The interference susceptibility of railroad signal equipment depends on the manufacturer, the type and model of equipment, as well as the frequency of the interference. Generally the interference tolerance of signaling equipment is not published by the manufacturer. However, the equipment manufacturers will often provide information on the expected compatibility of their equipment for specific anticipated or predicted rail-to-rail voltage environments. Track relays are generally quite insensitive to steady-state 60 Hz and harmonic voltages that are typically encountered in power line exposures. However, very low resistance relays, for example -ohm relays, have been found to be susceptible to interference effects for some operating conditions at the few-volt range. Highway grade crossing equipment (motion detection and constant warning time) tends to be the most susceptible of the commonly used signal equipment. Some units, generally at the lower operating frequencies, can have susceptibility threshold in the range of a volt at 60-Hz. Generally the susceptibility to 60-Hz decreases as the operating frequency increases, an interference tolerance of approximately 5 volts at 60 Hz is common for operating frequencies of 200-Hz and higher. The interference tolerance recommendation of AREMA is 5 volts81. Lower susceptibility to some power-line harmonic frequencies can occur for some operating frequencies. The equipment supplier can provide guidance on the expected susceptibility of specific grade crossing equipment models and operating frequencies. Some models can be obtained with an option that increases 60-Hz voltage tolerance beyond 5 volts. Steady State Coupling to Railroad Signal Lines Rail systems that have dc track signal systems must have a method for communicating trackcircuit occupancy information from an occupied track circuit to other track circuits that may be ahead or behind the train, in order to notify other trains of the occupied circuit or signal block. The occupancy information is generally transferred to other locations by pairs of signal wires, although radio relay links are now available. The most common method for transferring the occupancy information to other signal locations is by open-wire signal pole-lines that parallel the track. The information can also be sent to other signal locations by conductors of a signal cable, comprising multiple pairs of signal conductors. Voltage and current can be induced onto the signal conductors and the induced voltages may result in personnel safety concerns. Generally the signal conductors are physically close together so that only small voltage can be induced conductor-to-conductor. Therefore, the principal steady-state induced voltage concerns of signal wire conductors are generally related to personnel safety. Magnetic Field Coupling to Signal Lines The magnetic field of a power line that parallels a dc-signal railroad line can introduce voltage into an open-wire pole line or a signal cable. The magnetic field coupling to a signal wire circuit is similar to magnetic field coupling to the rail, except the signal wire circuit does not typically have losses such as are associated with the ballast under the rails. Thus, for a constant longitudinal electric field excitation, the voltage that is induced onto the signal conductors is nearly proportional to the length of the signal circuits. The length of signal-conductor circuits 12-71

Computer Modeling of Joint Corridors

can be longer than track circuits, since the conductors can extend to adjacent signal locations. Therefore, the magnetically induced voltage on signal-wire circuits can be higher than on track circuits. The magnetic field induced-voltage coupling to a signal cable depends on whether or not the cable is shielded, and how the shield is terminated to ground. If the shield is grounded at more than one location, current will flow in the cable shield as a result of the magnetic-field induction. The EPRI program CORRIDOR calculates the shield factor for cable shields that the user defines. The shield factor is the ratio of effective longitudinal electric field at a cable conductor with the shield to the field without the shield. A shield factor of zero is a perfect shield. The shield factor can also be expressed as the ratio of the voltage on the cable conductor with the shield to the voltage on the conductor without the shield. Even relatively modest shields can reduce the induced voltage on the core cable conductor to reduce the sheath-to-core touch potential to a safe value. The safe personnel touch voltage is generally considered to be higher for a line-wire signal conductor than for a rail conductor, because of the smaller contact area that is possible for the signal wire. The small diameter of the signal wire results in a higher skin-contact resistance to the wire than for larger conductors such as a rail. The higher skin-contact resistance helps to limit the current that can flow through a person that contacts an energized signal wire. The safe touch potential for signal-wire conductors is generally accepted as 50 volts [83]. Electric Induction Coupling to Signal Lines Concepts The power-system voltage on the power line conductors results in an electric field that is radial from each conductor. The lines of electric field terminate on the earth and are perpendicular to the surface of the earth. Therefore, near the earth surface, and in the region where railroad system conductors are generally found, the major component of the electric field is vertical, or perpendicular to the earth surface. The voltage difference between the power-line conductors and earth result in a small current that flows from each power conductor to earth. If another conductor, such as a railroad open-wire signal pole-line conductor is near the power line, some of the current that flows to earth from the power-line conductors will be intercepted by the signal conductor. Figure 12-55 shows a sketch of a power line with a nearby railroad open-wire signal pole line to illustrate the concept. The sketch shows that the power conductors are coupled to the signal wire line by distributed capacitance. The signal wire line also has a distributed capacitance to ground. The capacitances act similar to a voltage divider so that the voltage at a pole-mounted signal line conductor is a small fraction of the voltage on the power-line conductors. If the signal conductor is a part of a cable that is mounted on a pole, the conductors will also be at a similar potential. However, if the signal cable is buried in the earth, it will be shielded from electric induction coupling effects.

12-72

Computer Modeling of Joint Corridors

Figure 12-55 Power Line Voltage Coupling to Pole-Mounted Signal Conductor by Electric-Field Coupling

The current that flows to earth from the power line is called displacement current. The displacement current that is collected by an above ground open-wire signal pole line conductor normally flows to ground through the impedance to ground of the signal conductor. For signal pole-line conductors that are well maintained and mounted on low-leakage insulators, the impedance to ground is principally the capacitance to ground as is illustrated in Figure 12-55. The collected current that flows to ground through the capacitive impedance to ground of an open-wire signal pole line conductor causes a voltage drop across the impedance, which is the potential of the conductor, as measured with respect to ground. The conductor voltage can be quite high, because the impedance to ground of the conductor can be quite high. The voltage of a pole-line mounted conductor is also approximately equal to the ground-level vertical electric field of the power line, multiplied by the height of the conductor. These conditions can be summarized by:

Ev h = Vwire = I collected Z wire earth


where: Ev is the ground level vertical electric field, h is the mean height of the pole-line conductor,

Eq. 12-5

Vwire is the voltage of the pole-line conductor caused by electric field coupling, Icollected is the current induced into the pole-line conductor by the power line, per unit length of the exposure (per kft, per mile, etc.), 12-73

Computer Modeling of Joint Corridors

Zwire-earth is the impedance to ground per unit length of the exposure, typically the capacitive reactance per unit length of the conductor. As an example, Figure 12-56 shows the ground level vertical electric field for representative configurations of 525-kV transmission lines. The figure shows the ground-level vertical electric field in Volts/m versus distance from the center of the transmission line for three representative transmission-line configurations: flat (horizontal), vertical, and delta. At a representative distance from the transmission line, for example at 20-meters (65.6-feet), there is significant difference in the field intensity for the three different configuration transmission lines. The vertical electricfield values at that location are summarized in Table 12-3.
10 9 8 FLAT Electric Field at Ground - (kV/m) 7 6 5 4 3 DELTA 2 1 0 0 10 20 30 40 50 60 70 80 90 100 Distance from Line Center - (m)
10m 10.6m 10m 10.6m

VERTICAL 8.8 kV/m

V = 525 kV 3 X 3.3 cm CONDUCTOR, 45 cm SPACING (EQUIV. D = 30 cm)

10m 10m

10.6m

EQUILATERAL DELTA

FLAT

VERTICAL

Figure 12-56 Lateral Profiles of the Electric Field at Ground Level for Lines of Flat, Equilateral Delta, and Vertical Configurations [84] Table 12-3 Values of Field Strength at 20 meters (65.6 feet) from Three Configurations of 525 kV Transmission Lines Line Type Vertical Horizontal Delta Horizontal Vertical Field at Ground 0.3 kV/m 2.2 kV/m 4.9 kV/m Voltage (Space Potential) at 6m (20 feet) 1.8 kV 13.2 kV 29.4 kV

12-74

Computer Modeling of Joint Corridors

The table also shows the approximate voltage (space potential) for a wire height of 6 meters (approximately 20-ft.). The table and figure show that a significantly different voltage can result for different configurations of transmission lines. An important observation to make is that the voltages that are listed in Table 12-3 are all significantly higher than the steady-state voltage that is normally associated with personnel safety. The rail-to-ground voltage typically considered to be a limit for personnel safety is in the range of 25 volts. The fact that the electric field coupling to a signal pole line conductor may result in significantly higher voltage than is normally considered to be safe does not mean that these conductors are less safe. The confusion arises because of trying to use voltage as an indicator of safety, when the effect of electricity on the body is a result of the current that flows through the body. The internal impedance of the human body is typically in the range of 1,500 ohms, as is described in Chapter 7. The contact resistance, which is the resistance between the fingers or the hand and an electrified object, will add to the body resistance, as will the resistance that can flow from a person to earth through the feet contact, if the person stands on the earth. Thus limiting the current that can flow from an energized conductor through a 1500-ohm person to a safe value, independent of the conductor voltage is a better indicator than the voltage that may be measured, or calculated for an electrified conductor. Discussion associated with Equation 12-5 noted that the current per unit length that is collected by a signal pole line conductor is related to the voltage at the conductor location caused by the vertical electric field of the power line, and to the self-impedance to earth of the conductor, also per unit length (per kft, per mile, etc.). The total current that is collected by a conductor, for constant field excitation, is directly proportional to the length of the conductor, because the total impedance to earth of the conductor decreases with length. Figure 12-57 shows a person in contact with an insulated conductor that is excited by the vertical electric field. If the resistance of the person to earth is significantly less than the impedance of the conductor to ground, all or most of the current that is collected by the conductor will flow through the person. Thus, if the pole-line mounted conductor is relatively short, the current that is available to flow through a person may not be a safety concern, even though the voltage that is induced onto the conductor 3 by the electric field may be 1,000 volts or more . If the conductor is longer, the current that is available to flow through a person may be unsafe, even though the voltage that is induced onto the conductor by the electric field is the same.

It may not be possible to accurately measure the voltage on the conductor with a conventional voltmeter, because the impedance of the voltmeter may be too low, that is, not high enough with respect to the impedance of the conductor to earth.

12-75

Computer Modeling of Joint Corridors

Figure 12-57 Current to Person Touching Conductor that is Capacitively Coupled to Power Line

Electric Induction Coupling to Signal Lines Example This section presents analysis results for electric induction coupling of a 138-kV transmission line to a railroad system. Figure 12-58 shows a typical cross section within the exposure, which includes an open-wire signal pole line that is located on the opposite side of the track as the transmission line. Figure 12-59 shows the calculated space-potential profile, transverse to the transmission line, at a height representative of the signal pole line. The profile was calculated using the EPRI TLW (Transmission Line Workstation) program ACDC LINE module. The discussion above made note that an insulated conductor such as an open line-wire assumes the space potential at the location of the conductor. Also shown in Figure 12-59 is the voltage of a wire at the location of the signal pole line conductor as calculated by the computer program CORRIDOR. It is seen that the voltage predicted by the two programs at the location of the signal pole line is nearly the same.

12-76

Computer Modeling of Joint Corridors

Figure 12-58 Exposure of Rail System to 138-kV Transmission Line Typical Cross Section

12-77

Computer Modeling of Joint Corridors


7 Potential at Height = 20 ft 6 Potential at Height = 22 ft Open Circuit Wire Potential @ Height = 20 ft, Corridor

5 Space Potential (kV)

2 138 kV Location Track Location

Pole-Line Location

0 -100

-80

-60

-40

-20

20

40

60

80

100

Distance From Tower (feet)

Figure 12-59 Calculated Space Potential For 138-kV Transmission Line of Figure 12-58

The space potential profile is higher on the track side of the transmission line. The asymmetry of the space potential with respect to the center of the transmission line is dependent on the phasing of the transmission line conductor voltages. The figure shows that the space potential (or the ideal open-circuit voltage) of a conductor at the location of signal pole line is approximately 1000 volts. These calculations were made for an equivalent pole-line circuit pair of conductors (two conductors spaced by 11.5 inches) and the shielding effect of other pairs was ignored. Other conductors on the signal pole-line can reduce the conductor voltage, as can insulator leakage, leakage resistance and capacitance of terminating signal equipment, nearby trees or other close objects that have some conductivity to earth. Of more practical importance for personnel safety, as discussed earlier, is the current that can be drawn by a person who touches the conductor excited by electric induction, as is depicted in Figure 12-57. Figure 12-60 illustrates the predicted voltage for a conductor at the signal pole-line location, and the current that would flow through a 1500-ohm person touching the conductor. The analysis for the figure assumes excitation of a signal pole-line conductor by a 138-kV transmission line as shown in Figure 12-58, with the length of the exposed conductor as a variable. The predicted open circuit voltage of the conductor is displayed on the left axis and the current through the person is on the right axis of the figure. It is seen that the open circuit voltage of the conductor is not dependent on the conductor length, but the current to earth through a person is directly proportional to the conductor length. For the assumed geometry, 12-78

Computer Modeling of Joint Corridors

the steady state current through the person is higher than 10 milliamperes for a conductor length in excess of approximately 10,000 feet.

Figure 12-60 Electric Induction Voltage on Isolated Wire and Current to 1500-Ohm Person versus Excited Length of Wire

The electrical resistance of a person is an important variable that has an influence on the current that will flow when an energized conductor is touched. The 1500 ohms assumed for the above example is representative of the total body impedance for a hand-to-hand or hand-to-feet current path with large area contact, for approximately 95% of the population [85]. For smaller contact areas, such as contacting a wire, the contact impedance may be significantly higher. For low source impedance circuits that are contacted, the skin resistance may significantly limit the body current. However, a well insulated electric induction excited wire, such as a signal pole-line conductor can be a high impedance source of current. The current that can flow through a person from a high impedance source of current can be rather insensitive to the impedance of the person. Figure 12-61 illustrates the effect of the total body impedance of a person on the electrically induced current that will flow from a wire through the person in contact with the wire. For this plot, the geometry of Figure 12-58 was assumed, with the 138-kV line at a distance of 60 feet from a 10,000-foot length of an isolated pair of pole-line conductors. The current through the 12-79

Computer Modeling of Joint Corridors

simulated person impedance is plotted versus the impedance of the person. It is seen that the current is maximum for a low-impedance person, but only decreases by approximately 20% for a body impedance of 75,000 ohms. It is clear from Figure 12-61 that the electrostatically-excited isolated conductor acts as a current source and that nominal insulation of the person by standing on gravel may not prevent the current from flowing through the person. The foregoing paragraphs have presented an overview of important characteristics of the exposure geometry of Figure 12-58. The following paragraphs will show calculated results for specific signal line-wire circuits that exist within an existing exposure region.
12

10,000 foot Length Exposed Conductor


10

Current Through Person (mA)

Electrostatic Excitation

4
DISTRIBUTED CURRENT

0 0 20,000 40,000 60,000 80,000 100,000 120,000

Resistance of Person to Earth (ohms)

Figure 12-61 Effect of Total Body Impedance on Shock Current from Electric Induction Excited Isolated Conductor

The analysis of the electric induction coupling to the signal pole-line modeled three circuit pairs as individual conductors. The east-bound distant circuit lengths provided by the railroad were used for assessing the electrically induced current that would flow to a person that contacted one of the conductors of the circuit. Table 12-4 identifies the milepost (MP) end points and the lengths for those circuits. For each circuit, the circuit pair of wires was terminated at one end of the circuit by a battery and resistor, and by a high-impedance relay at the other end. A distributed leakage resistance of 1 megohmmile was assumed for each of the pole-line conductors to account for leakage of the mounting insulators. In some regions the signal conductors are in a buried cable rather than mounted on the pole line. The buried cable shields the conductors from the electric field in those regions. In addition to the east and west-bound distant circuits, an additional pair of conductors was considered to be very long and continuous to simulate the code line. 12-80

Computer Modeling of Joint Corridors Table 12-4 Signal Pole-Line Circuits Used to Evaluate Available Electric Induction Coupled Shock Current From Signal 83 (MP 83.1) Signal 86 (MP 86.36) Signal 89 (MP 89.62) CP 92 (MP 92.91) CP 93 (MP 93.88) Signal 96 (MP 96.91) To Signal 86 (MP 86.4) Signal 89 (MP 89.62) CP 92 (MP 92.91) CP 93 (MP 93.88) Signal 96 (MP 96.91) Signal 99 (MP 99.29) Length (miles) 3.26 3.26 3.29 0.97 3.03 2.38

Figure 12-62 shows the predicted electrically-induced current through a 1500-ohm resistance connected between an east-bound distant (EBD) circuit conductor and ground. This resistance simulates a grounded person touching the conductor. The simulated person resistance was moved sequentially from one circuit to the next to obtain the current values shown in the figure. Two curves are shown in the figure:

the lower curve includes the shielding effects of the long code line, the effect of the code line was minimized for the upper curve by segmenting the code line conductors into smaller lengths.
25

20 Current Through 1500 Ohm Man

15

10

With Code Line Effect Without Code Line Effects

0 102

100

98

96

94

92

90

88

86

84

82

Railroad Milepost

Figure 12-62 Predicted Current to Earth Through 1500 Ohm Person From Electric Induction Excited Pole Line Signal Circuit

12-81

Computer Modeling of Joint Corridors

Since the code line is long and extends well beyond the exposure region, its voltage is low because its impedance to earth is low as a result of its long length. The presence of the code line, or other long conductors on the pole line, changes the field distribution near the pole line and reduces the voltage and current collected by the other shorter conductors. Figure 12-63 compares the predicted open circuit voltage to earth of the EBD circuit conductors with and without the shielding effect of the long code line. It is seen that the presence of the code line reduces the predicted voltage of the other conductors.
1200

1000

Open Circuit Voltage

800

600

400

200

With Code Line Effect Without Code Line Effect

0 102

100

98

96

94

92

90

88

86

84

82

Railroad Milepost

Figure 12-63 Predicted Open Circuit Voltage of Electric Induction Excited Signal Pole Line Circuits

The current predicted to flow through a person in Figure 12-62 is in the range of several milliamperes for most of the existing distant pole-line circuits within the exposure region. The physiological effect of current through a person becomes an important consideration for current values that exceed the perception threshold, which is typically in the range of 0.5 to 1.0 mA. Significant involuntary reactions have been reported for current in the range 2.2 to 3.2 mA [86]. At higher values of current, a person may not be able to let go of a grasped conductor. Information on the inability to let go of a grasped conductor as a result of 60 Hz current from the conductor through the arm generally references the work of Dalziel. He summarizes the results as a 99.5 let-go percentile (that is, 99.5 percent of test subjects could let 12-82

Computer Modeling of Joint Corridors

go of the conductor at the stated current) of 9 mA for men and 6 mA for women, with mean values of 15.87 mA for men and 10.5 mA for women [87]. The IEC recommends an average value of 10 mA for long-duration current (greater than 2 seconds) for let go [85]. For current that is less than the let-go current, there usually are no harmful physiological effects. For current that is higher than the let-go value, there is a likelihood of cramp-like muscular contractions and difficulty in breathing for current durations longer than two seconds. When the electrically-induced current onto open-wire signal pole-line conductors exceeds safe touch potential, the pole-line is typically removed, although the length of circuits might be reduced by using repeater relays, which may reduce the current that can flow through a person who contacts the circuit. The added repeater increases the maintenance of the signal pole line and is therefore not a desirable mitigation for the railroad that operates the signal line. Removing the signal pole-line is also a consideration, and there are several methods to provide the necessary signaling without the open-wire signal pole-line. (The presence or absence of the open-wire signal pole-line does not have any significant influence on the magnetic-field coupling to the track system.) Replacing the signal pole-line with a buried cable is an option. This would circumvent the problem of electric induction coupling to the pole line, but other approaches are generally preferred by the railroads. The high installation and maintenance costs of buried signal cable make other alternatives preferable. The railroads are systematically replacing existing dcsignal systems, which have open-wire pole-lines, with other newer technology signal systems. Typically the dc-signal systems are replaced with coded electronic track-circuit systems. The dc track signal system can be retained without a hardwire (pole-line or cable) link between signal locations by using wayside spread-spectrum radio systems. Power Line Fault Coupling to Railroad System Overview A line-to-earth fault along a transmission line results in a high current in the faulted phase, which can be considerably higher than the normal load current. The high current only exists for a brief period, since sensing circuits on the power system will open breakers to stop the flow of current. The high fault current in the transmission line conductor creates a high magnetic field, which can induce high voltage and current in long parallel conductors such as a rail system or pipeline. If a line-to-structure fault occurs within the exposure, high current can flow to earth at the faulted structure and other nearby transmission line structures that are connected together by the shield wire. The earth current from a faulted structure can flow to the rails or increase the rail-to-earth potential in the vicinity of the structure. The railroad signal equipment is protected against high-current events such as lightning by lightning arresters that are installed at signal-equipment locations. These lightning arresters will fire if the fault-induced rail-to-ground voltage exceeds the spark-over potential of the arresters. The fired arresters will protect the railroad signaling equipment from damage during the short 2 duration of the fault, if the arrester (i t) rating is not exceeded by the fault-induced current that flows through the arrester. Thus one of the conditions of concern for the faulted condition is to evaluate the current that may flow through fired railroad signal-system lightning arresters relative to the their rated current. 12-83

Computer Modeling of Joint Corridors

The firing of track lightning arresters results in the rails being connected together and to the primary earth connection at signal-equipment enclosure locations by the low-impedance actively firing arresters. The arresters protect not only the equipment, but personnel that may be in contact with any of the circuitry within an equipment enclosure, since the enclosure is also connected to the primary ground at the signal-equipment location. However, a person who is outside of the equipment enclosure, but in contact with it, or who contacts a rail, a track-lead, or a rail car may experience a high touch potential relative to local earth if the rail-to-soil touch potential is higher than a safe-touch value. Fault-Induced Firing of Track Arresters The magnetic field from the current in the faulted transmission line will produce a voltage in the exposed track. The magnitude of the voltage and the current that flows in the track will depend on whether or not the lightning arresters associated with the track actually fire, and the extent of track over which such firing occurs. Rail-to-ground and rail-to-rail lightning arresters are typically located at the inputs to all track signal equipment. Therefore, rail-to-ground arresters are located at each insulated joint and at locations between IJs where signal equipment is located. For this discussion, we are primarily concerned about the arresters that are connected rail-to-ground on each side of track insulated joints. Figure 12-64 is a sketch that shows lightning arresters that are connected from rail-to-ground on each side of an insulated joint. The fault-induced voltage Ve and Vw on each side of an insulated joint depends on numerous factors, including:

the fault-current magnitude, the length of each track circuit, the track circuit ballast resistivity, the soil resistivity, the separation between track and fault current.

The voltage at which rail-to-ground lightning arresters fire, may vary over a significant voltage range. The 60-Hz breakdown voltage of air-gap arresters are quoted in manufacturers literature as ranging from 350 to 950 volts. Gas tube arresters typically fire at lower voltage.

12-84

Computer Modeling of Joint Corridors

Figure 12-64 Fault-Induced Voltage Across Track Lightning Arresters

The Firing of Arresters in One Track Circuit

Once the arresters fire at one or two locations, the rail-to-ground voltage generally increases at the opposite ends of a track circuit. This is illustrated in Figure 12-65a, which shows the rail-tosoil voltage profile along a track circuit that could occur as a result of fault-related magneticfield induction, if no arresters fired. The variability in firing potentials of typical track lightning arresters may result in the firing potential being exceeded at one end of a track circuit before it is at the other end. In Figure 12-65a, if the voltage Ve exceeds the arrester firing potential, but the voltage Vw does not, the rail will be connected to earth at the Ve end by the low impedance of the fired arrester. When that happens, the voltage profile along the rail will tend to appear as is shown in Figure 12-65b. In Figure 12-65b the voltage Ve is nearly zero, while the voltage Vw becomes approximately twice as high as it was in the (a) sketch. Thus, the firing of the arrester at one end of the track circuit causes the voltage at the opposite end of the track circuit to increase. The increased voltage across the arrester at the opposite end of the circuit may be sufficient to also break down that arrester.

12-85

Computer Modeling of Joint Corridors

Figure 12-65 Fault-Induced Rail Voltage Profiles with and Without a Fired Arrester

The Firing of Arresters in Adjacent Track Circuits

The firing of the rail-to-ground lightning arresters in one track circuit by induced voltage caused by magnetic field coupling from a transmission line fault will increase the voltage across the unfired arresters in adjacent track circuits. Figure 12-66 is a schematic representation of one rail at an insulated joint that has one fired arrester and one non-fired arrester. The fired arrester on 12-86

Computer Modeling of Joint Corridors

the right side of the IJ is represented as a low-value resistance Ra, and the grounding resistance for the arresters Rg is also shown. If both arresters in the track circuit on the right side of the IJ fire (for the reasons discussed with Figure 12-65), current will flow through the fired arrester and through the grounding resistance Rg. If the voltages are the result of magnetically-induced voltage, and several track circuits are similarly excited, the current that flows through the fired arrester and the grounding resistance will cause a higher voltage Va1 across the unfired arrester than if the arrester on the right side of the IJ were not fired. The voltage across the arrester on the left side of the IJ, without the fired arrester on the right side of the IJ, is approximately (Va1 = Vr1 ) , where Vr1 is the rail-to-remote (approximately the same as the rail-to-ground) voltage. However, when the arrester on the right side of the IJ is fired, the voltage across the unfired arrester is (Va1 = Vr1 + Vg ), which is a higher voltage, because of the current that flows through Rg and Ra. The voltage Vg is dependent on the grounding resistance Rg but will tend to be in the same range as the rail-to-ground voltage that would exist on the right side of the IJ, if the arrester on the right side of the IJ were not fired.

Figure 12-66 Fault-Induced Arrester Voltages

The discussions associated with Figure 12-65 and Figure 12-66 illustrate:

if fault induced voltage fires a rail-to-ground arrester on one end of a track circuit, the voltage across the arrester on the opposite end of the track circuit will increase, increasing the chance that that arrester will also fire, the firing of arresters on both ends of a track circuit results in current flow through the arrester grounding connections in that circuit, 12-87

Computer Modeling of Joint Corridors

the current flow through the fired arresters and grounding connections causes additional voltage across the arresters in the adjacent track circuits, which may cause those arresters to also fire.

The high magnetic field induction produced by a power-line fault is generally assumed to cause the track lightning arresters to fire in a cascade throughout the common-corridor region, and beyond. The firing of the track arresters over a significant segment of a long exposure has the effect of making the rails nearly continuous over the extent of the exposure, which can result in substantially higher rail current (which flows through fired arresters) than if the arresters were not fired in an extended region. The following paragraphs illustrate an analysis of fault-induced cascade firing of track lightning arresters as a sequence of steps for a specific multi-block exposure that has been studied. For the analysis, we assumed that those arresters with more than 1000 volts would break down. If more than 1000 volts was predicted across an arrester the arrester was replaced by a low impedance and connected to a 15 ohm earthing resistance (to simulate the equipment housing ground) in the model for use in the next iteration. Figure 12-67 shows the predicted induced voltage across each rail-to-ground arrester, at all IJs that were not bypassed by wide-band shunts for several iterations of the analysis. Figure 12-67 (stage 1) is for the condition that none of the arresters have fired. At each arrester location, the voltage across the arrester on the west side of the IJ is shown by the darker bar, which is in front of the east-side arrester voltage bar. For the analysis, we next assumed that those arresters with more than 1000 volts in Figure 12-67 (stage 1) would break down; those arresters were replaced in the model by a low impedance and connected to a 15 ohm equipment-enclosure ground. If only one rail arrester fires at an IJ, the voltage across the arrester on the opposite side of the IJ increases because of the firing of the first arrester, as described with Figure 12-66. If both arresters fire, the effective continuous length of the rail increases, which increases the voltage at subsequent IJs. Figure 12-67 (stage 2) shows the predicted voltages at the arrester locations after the second iteration. It is seen in Figure 12-67 (stage 2) that the voltage exceeds 1000 volts at several arrester locations. Those arresters all had less than 1000 volts for the first iteration of Figure 12-67 (stage 1). Thus, for the third iteration, the arresters that exceeded 1000 volts in Figure 12-67 (stage 2) were assumed to fire, resulting in the predicted voltages of Figure 12-67 (stage 3). In Figure 12-67 (stage 3) it is seen that all the arresters, except one near each end of the exposure are fired. However, the voltage across those two arresters exceeds 1000 volts, and therefore they will also fire. Subsequent iterations are shown in Figure 12-67 (stage 4) and Figure 12-67 (stage 5). In Figure 12-67 (stage 5), all the arresters within the exposure except one near each end are fired. Those arresters, at approximately MP81 and MP 101 are at less than 1000 volts, and may not fire. However, the results of Figure 12-67 indicates that rail-to-ground lightning arresters within an extend region of exposure can be expected to fire as a result of a phase-to-ground fault on a paralleling transmission line.

12-88

Computer Modeling of Joint Corridors

Figure 12-67 Fault-Induced Sequence of Track Arrester Firing Along Corridor

The firing of lightning arresters by fault-induced current is not detrimental. The fired arresters help protect the track-connected equipment from the short-duration overvoltage that can be caused by a power transmission line fault. The above sequential fault-induction modeling of the exposure included the effect of the grounding resistance of equipment primary grounds located at normally non-bypassed IJs. The analysis did not include connection to the rails of primary grounds at other equipment locations, which would occur because the arresters at those (non-IJ) locations may also fire. The effect of grounding resistance at those other equipment locations is similar to a lower ballast resistivity and would result in lower rail voltage, which would possibly limit the track length of arrester firing to somewhat less than for this analysis.

12-89

Computer Modeling of Joint Corridors

The firing of the track arresters over a significant segment of a long exposure has the effect of making the rails nearly continuous over the extent of the exposure, which can result in substantially higher rail current (which flows through fired arresters) than if the arresters were not fired. The following section discusses the rail current that is expected to flow as a result of track arrester firing throughout an exposure region. Fault-Induced Rail Current The induced rail current for fault condition is important because that current must flow through the fired track lightning arresters at track insulator locations. Figure 12-68 illustrates the current at a rail insulated joint with fired rail-to-ground lightning arresters on each side of the IJ. The principal path for the rail current Ir is through the two fired arresters from one rail segment to the next, around the IJ, with a smaller current, Ie , that flows to earth through the signal-equipment grounding resistance. The induced rail voltage can vary along the exposure. Therefore, even if the grounding resistance at the signal-equipment enclosures is constant, the current through the earth connection Ie and the current that flows in the rails, around the IJ through the fired arresters will vary over the exposure.

Figure 12-68 Fault-Induced Current at Fired Arresters

The current that flows through the fired arresters is very important, since the current must remain within the rated current limits of the arresters. Specific information on the power fault performance of railroad track lightning arresters is very limited. Information on the 60-Hz faultcurrent rating for these devices is generally not available from the manufacturer. The current carrying capacity of an arrester depends on the duration of the waveform. AREMA performance guidelines are for lightning-type waveforms, which have relatively short duration compared to the duration of a power line fault-induced current waveform. The AREMA guideline for line-toground vital signal track circuit primary surge protectors must be capable of discharging 20,000 amperes with an 8x20 microsecond waveform [88]. A common method of characterizing the performance of surge protectors is by the I2t (Joules/ohm) rating. The I2t value of the AREMA 12-90

Computer Modeling of Joint Corridors

lightning waveform is approximately 8,000 Joules/ohm. An equivalent I2t value for a 0.1 second fault-current waveform has a current value of 280 amperes. Limited but well documented tests performed by Ontario Hydro were published by the AAR in 1984 [89]. The tests used a 60-Hz current duration of 12-cycles for current up to 1000 amperes. Some arresters failed these tests by consistently becoming shorted for current between 250 A (I2t = 12,500) and 1000 A. However for those tests, some other types of arrester 2 survived 12-cycle faults up to 1000-amperes (I t = 200,000), which for a 6-cycle fault is approximately 1,400-amperes, based on equivalency of I2t. The above section illustrates that for long exposures the fault-induced rail-to-ground voltage tends to fire the track lightning protector over a significant portion of the exposure (and possibly beyond). The fault-induced current that flows through the fired arresters tends to be higher for a long exposure, since the effects of the ends of the exposure are reduced for long exposures. For sufficiently long exposures the fault current that flows through fired arresters is relatively independent of the track ballast resistivity, but is higher for high values of soil resistivity. The induced rail current will be higher for a very long uniformly excited exposure than for shorter exposures. Figure 12-69 shows the predicted fault-induced rail current for a long exposure by a 115-kV transmission line. The curves relate the induced rail current (amperes/kA fault current) versus separation of the track from the power line. The induced rail current is shown for two values of soil resistivity. The results will be similar for other transmission configurations and voltage classes. The results of Figure 12-69 show that arresters that would tolerate 300 amps/kA fault current would be suitable for a long exposure. To satisfy that condition for arresters that just meet the AREMA, guidelines the fault current would have to be less than 1000 amps, if the fault clearing time is 0.1-second (a six-cycle fault). However, if the arresters could reliably dissipate the energy with I2t = 200,000, as suggested by the testing of Reference [89], a six-cycle fault could be approximately 4.6-kA.

12-91

Computer Modeling of Joint Corridors


500

Rail Arrester Current (amps/kA fault current)

400

115 kV Vertical Line with 3/8 HSS OHSW

300

200

Soil Resistivity Ohm-m 1000 100

100 Rail Current

0 10 100 1000 10000


Distance Power to Track (feet)

Figure 12-69 Calculated Fault-Induced Rail Current for Long Exposure

Although the predicted maximum induced track current, for fired track arresters and a very long exposure is not dependent on the ballast resistivity, the induced current for typical exposures does depend on the track ballast resistivity. Typically, the predicted induced rail current is higher for lower values of track ballast. Therefore, low values of track ballast should be assumed to assess the fault-induced rail current. The investigation of transmission-line fault-induced effects should take into consideration the variation of fault current that can occur along the length of the exposure. If the exposure length is several miles, there can be significant variation in fault current for faults at different locations. A fault can occur at almost any location along a transmission line. The fault current that is fed to a fault from a substation decreases as the distance to the fault increases, because the transmission line impedance between the substation and the fault location is higher for greater distance to the fault. Typically, a transmission line connects between two substations, and current is supplied from both substations to the fault. Figure 12-70 shows the predicted single-phase to ground fault current that is fed from each direction to a fault along railroad exposure. Since the location of a fault is unknown, analysis of coupling to the railroad system should be made as function of fault location.

12-92

Computer Modeling of Joint Corridors


16,000

14,000

12,000 Current to Fault (amperes) From Downstream Source From Upstream Source 8,000

10,000

6,000

4,000

2,000

0 15 16 17 18 19 20 21 22 23 24 Railroad Milepost

Figure 12-70 Single-Phase to Ground Fault Current vs. Fault Location

Figure 12-71 shows the calculated rail-current profile for several locations of a phase-to-tower fault for an exposure that has been investigated for low (10-ohmkft) track ballast. Each curve shows the maximum current in the four main-line rails at each location. The envelope of the composite curve-set defines the fault-induced current that can flow in the rails, since a fault can occur at any location within or beyond the exposure region. Figure 12-72 shows a similar curve set for assumed 100-ohmkft ballast. The curves of these two figures show that the induced current is higher for the lower ballast-resistivity condition, which is typical. The ballast may vary over this range depending on the wet-dry cycle. Therefore, analysis to determine the faultinduced current that is likely to flow through fired track arresters should be performed as a function of fault location, at sufficiently close spacing to permit assessment of the maximum arrester current that should be anticipated.

12-93

Computer Modeling of Joint Corridors


1200

1000 Rail Current (60-Hz rms amperes)

Assumed: Earth Resistivity = 100 ohm-m Ballast Resistivity = 10 ohm-kft Track Arresters Fired in Exposure Region

Fault Location 800 MP 16.7 MP 17.82 MP 18.86 MP 19.82 MP 20.84 MP 21.87 MP 22.8

600

400

200

0 0 5 10 15 Railroad Milepost 20 25 30

Figure 12-71 Calculated Fault-Induced Rail Current for Several Fault Locations
700 Assumed: Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Track Arresters Fired in Exposure Region

600

Rail Current (60-Hz rms amperes)

500

Fault Location MP 16.7 MP 17.82 MP 18.86 MP 19.82 MP 20.84 MP 21.87 MP 22.8

400

300

200

100

0 0 5 10 15 Railroad Milepost 20 25 30

Figure 12-72 Calculated Fault-Induced Rail Current for Several Fault Locations

12-94

Computer Modeling of Joint Corridors

The curves show that current in the range of 1000 amperes can be expected to flow in the rails at some locations and thus through the rail-to-ground lightning arresters if they exist at those locations. The model used to analyze the rail current is a small-signal or linear model. The literature indicates that the impedance of rails increases somewhat at higher rail currents, such as are predicted in the above figures [90, 91]. We adjusted the internal rail impedance in the model to correspond to the reported measured impedance for currents in the range of 1000 amperes. That adjustment resulted in maximum currents that were approximately 12% less than are obtained with the linear model. Thus, the fault-induced rail current that is calculated using a linear model is expected to provide a conservative estimate of fault-induced rail current. Fault-Induced Rail Voltage and Personnel Safety The high current in the faulted phase results in induced rail-to-earth voltage that may exceed a safe touch potential for the short-duration period before the transmission line circuit breakers de-energize the transmission line. The most-likely location for railroad signal maintenance personnel to encounter a fault-induced high touch potential is may be at signal-equipment locations along an exposure, however, at other locations track maintenance personnel may make contact to the rails. Therefore, evaluation of the fault-induced track voltage is generally performed as a part of a compatibility investigation. If the induced rail voltage exceeds the breakdown of lightning protectors at signal-equipment locations, the arresters will fire. The fired arresters connect the rails to the primary ground-rod and thus to earth at signal-equipment housing locations. The fired arresters protect not only the equipment, but personnel that may be in contact with any of the circuitry within an equipment enclosure, since the enclosure is also connected to the primary ground at the signal-equipment location. However, a person who is outside of the equipment enclosure, may be in contact with the enclosure, the rails, or any other signal conductors. He may experience a high touch potential to those items relative to local earth, because of the fault-current coupling to the rail system. A recognized guideline for determining a safe touch potential for conductors that may be energized to a high voltage for a short period of time is the IEEE Std. 80 [92], which is based on the work of Dalziel. The guideline procedure permits safe touch potentials to be calculated based on:

a safe withstand body current, which depends on

the size of the person the square root of the fault duration.
the resistance of the body, the resistance of the feet contact to earth.

12-95

Computer Modeling of Joint Corridors

The AAR Principles and Practices For Inductive Coordination Of Electrical Power And Railroad Communication/Signal Systems the Blue Book [93], identifies tolerable faultcondition induced voltage objectives for personnel safety as 430 V rms for the usual power line equipment and maintenance, or 650 V rms for high reliability power lines with high speed relaying and fault clearing. The AAR Blue Book limits do not quantify the fault clearing time. But it is clear that that the two voltage values identified in the Blue Book depend on the (undefined) fault duration. The values given in the Blue Book do not depend on the size of the person, which is common in the European guidelines such as the CCITT Directives. Figure 12-73 shows the IEEE Std 80 guideline safe touch potential values for a 70-kg (155-lb) person. The safe touch potential assumes hand-to-feet current flow through a person standing on soil, with no shoe insulation included. Three values of soil resistivity are shown in the figure. The resistivity of the soil affects the flow of current from feet to earth for the IEEE calculation method. The IEEE-based curves on the figure give the touch potential as a function of the duration of the current waveform (the fault-clearing time). The safe-touch potential guidelines of the 1977 revision of the AAR Blue Book are also shown on the figure. The IEEE Std 80 notes that the 70-kg person is suitable for workers within power substations.

10000

70 kg (155 lb) person


Safe Touch Potential (volts) Soil Resistivity (ohm-m)

1000 500
1000

IEEE Std 80 Guideline for 99.5% Safe Touch Potential

100
High Reliability Power Usual Power Equipment AAR 'Blue-Book'

100 0.01

0.1

Shock Duration (seconds)

Figure 12-73 Fault-Induced Touch-Potential Safety Guidelines (70 kg person)

The IEEE Std 80 also notes that outside the substation, where the general public may be a concern, an assessment should be made on the use of a 50-kg (110-lb) safe-touch potential criterion. Figure 12-74 shows IEEE-based safe-touch potential curves for a 50-kg person. Since railroad tracks are accessible to the general public, consideration may be needed on establishing appropriate limits for fault-induced rail potentials. Special situations have arisen where smaller 12-96

Computer Modeling of Joint Corridors

sized persons may require consideration, such as at stations where people board and disembark from trains that are on tracks that may be affected by fault-induced potentials. Additional discussion of the personnel-safety topic is presented in Chapter 7.

10000

50 kg (110 lb) person

Safe Touch Potential (volts)

Soil Resistivity (ohm-m)


1000 500 1000 100

AAR 'Blue-Book' High Reliability Power

Usual Power Equipment

IEEE Std 80 Guideline for 99.5% Safe Touch Potential


100 0.01 0.1 1

Shock Duration (seconds)

Figure 12-74 Fault-Induced Touch-Potential Safety Guidelines (50 kg person)

The safe-touch potential for a person standing in the ballast region near the tracks may be higher than for standing on the native soil, as a result of the added resistance provided by the layer of stones in the ballast. The above figure show that the safe touch potential increases as the soil resistivity increases. The voltage, which is induced onto railroad tracks that parallel a transmission line, also increases with soil resistivity. Often, the resistivity of the soil is not well known, and analysis is made for a representative value of soil resistivity. A value of soil resistivity that is often provided by power companies as a representative value to use for initial investigations is in the range of 100-ohmm. The question arises as to how changes in soil resistivity relative to a representative value might affect whether or not an induced voltage is safe for other values of soil resistivity. Figure 12-75 summarizes the analysis results of how the induced voltage and the safe touch potential vary as the soil resistivity (assumed to be uniform with depth) deviates from a reference value of 100-ohmm. . The horizontal axis is the soil resistivity. The vertical axis is the ratio of the safe touch potential Vsafe touch to Vrail-to-ground , each normalized by its value at 100-ohmm soil 12-97

Computer Modeling of Joint Corridors

resistivity. The ratio of the Vs is plotted as a function of soil resistivity. It is seen that the plot has a minimum at a soil resistivity of approximately 100-ohmm.
1.8

1.7

1.6 Ratio (Vsafe touch / Vrail-to-ground )

1.5

1.4

1.3

1.2

1.1

1 10 100 Soil Resistivity () ( ohmm) 1000

Figure 12-75 Relation Between Safe Touch Potential and Induced Voltage vs. Uniform Soil Resistivity

The plot of Figure 12-75 shows that for soil-resistivity values greater than 100-ohmm, the safe touch potential increases faster than the induced voltage, and for soil-resistivity values smaller than 100-ohmm, the touch potential decreases slower than the induced voltage. Therefore, if the actual soil has higher uniform resistivity than 100-ohmm, the prediction results in a higher voltage than for the actual soil resistivity (a conservative prediction). Similarly, if the actual soil has lower uniform resistivity than 100-ohmm, the prediction results in a higher voltage than for the actual soil resistivity, which is also conservative. Thus, the use of 100-ohmm soil resistivity for predictive analysis appears to be an approximate worst case value, which provides a conservative estimate of touch potential if the actual soil resistivity is not known. Figure 12-76 shows an example of calculated rail-to-earth voltage profile for several locations of a phase-to-ground fault along the same exposure for which rail-current were presented in Figures 12-71 and 12-72 for low (10-ohmkft) track ballast resistivity. Each curve shows the maximum voltage in the four main-line rails at each location along the profile. Figure 12-77 shows a similar curve set for assumed 100-ohmkft ballast. The curves show that the induced voltage is higher for the high ballast-resistivity condition. The ballast may vary over this range depending on the wet-dry cycle, therefore it is prudent to perform the analysis for a ballast resistivity that results in near worst-case induced voltage. Generally, the induced rail voltage does not increase significantly for ballast resistivity that is higher than about 100-ohmkft, so that value appears to be a reasonable choice for assessing touch-potential values that may be anticipated. The envelope of the curve-set that is obtained for faults at progressive locations along the exposure can be used to define a composite fault-induced touch voltage. Since a fault can occur at any location within or beyond the exposure region, the envelope of the maximum induced voltages as 12-98

Computer Modeling of Joint Corridors

a function of fault location, such as is shown in Figure 12-77, can be used to assess the safety for persons who may contact the rails or rail-cars or other objects that are at the same potential as the rails with fired lightning arresters to ground. Such other objects include the signal-equipment enclosures, which are connected to the rails through fired arresters.
2,000 1,800 1,600 1,400 1,200 1,000 800 600 400 200 0 0 5 10 15
Railroad Milepost

Assumed: Earth Resistivity = 100 ohm-m Ballast Resistivity = 10 ohm-kft Track Arresters Fired in Exposure Region

Rail-to-Earth Voltage (volts)

Fault Location MP 16.7 MP 17.82 MP 18.86 MP 19.82 MP 20.84 MP 21.87 MP 22.8

20

25

30

Figure 12-76 Calculated Fault-Induced Rail-to-Earth Voltage for Several Fault Locations

12-99

Computer Modeling of Joint Corridors


5,000 4,500 4,000 3,500 3,000 2,500 2,000 1,500 1,000 500 0 0 5 10 15
Railroad Milepost

Assumed: Earth Resistivity = 100 ohm-m Ballast Resistivity = 100 ohm-kft Track Arresters Fired in Exposure Region

Rail-to-Earth Voltage (volts)

Fault Location MP 16.7 MP 17.82 MP 18.86 MP 19.82 MP 20.84 MP 21.87 MP 22.8

20

25

30

Figure 12-77 Calculated Fault-Induced Rail-to-Earth Voltage for Several Fault Locations

Simple metallic grounding mats or grids that are placed at signal-case locations and connected to the signal-system ground can provide safe touch potentials at those locations. Spreading a layer of high-resistivity crushed stone (granite) past the touch distance from equipment enclosures can also aid in increasing the safe touch potential at these locations.

References
76. J. R. Carson, Wave Propagation in Overhead Wires with Ground Return. Bell System Technical Journal, Vol. 5, 1926, p.539-554. 77. J. R. Wait and K. P Spies. On the Image Representation of the Quasi-Static Fields of a Line Current Source Above the Ground, Canadian Journal of Physics, vol. 47, 1969, pp. 2732-2733. 78. R. G. Olsen and T. A. Pankaskie, On the Exact and Image Theories for Wires at or Above the Earths Interface, Paper 82 SM 391-1, IEEE Power Engineering Society 1982 Summer Meeting, July 1982. 79. Cramer, B. S. Phase Imbalances on ComEds Transmission and Distribution Systems, Proceedings of the American Power Conference, April 1995. 80. AREMA Communications & Signal Manual Parts 3.1.20 and 3.1.26. 12-100

Computer Modeling of Joint Corridors

81. AREMA Communications & Signal Manual Part 11.4.1. 82. PSC Staff Report: Wisconsins Stray Voltage Experience An Update, Public Service Commission of Wisconsin, April 1998. 83. Principles and Practices for Inductive Coordination of Electric Supply and Railroad Communication/Signal Systems, Association of American Railroads, September 1977. 84. Transmission Line Reference Book, 345-kV and Above, Second Edition, 1982, Electric Power Research Institute. 85. International Electrical Commission Report IEC 479-1, Effects of Current on Human Beings and Livestock, 1994. 86. Patrick Reilly, Electrical Stimulation and Electropathology, Cambridge University Press 1992. 87. C. F. Dalziel, Reevaluation of lethal electric currents. IEEE Transactions Industrial Applications IGA-4(5), 1968, 467-476. 88. 60-Hz Fault Current Withstand Tests on Signal Lightning Arresters AAR C&S Technical Paper 1984 89. AREMA Communications & Signal Manual Part 11.3.1. Recommended Function of Primary Surge Protectors for Electrical Surge Protection of Signal Systems. 90. Modeling of Nonlinear Rail Impedance in AC Traction Power Systems, Hill, R.J. and Carpenter, D.C., IEEE Transactions on Power Delivery, Vol. 6. N0. 4, October 1991. 91. Investigation of Rail Impedances, Trueblood, H.M. and Wascheck, G., Electrical Engineering, December, 1933. 92. ANSI/IEEE Std 80, IEEE Guide For Safety in AC Substations Grounding, The Institute of Electrical and Electronics Engineers, Inc., 345 East 47th Street, New York, NY 10017. 93. Principles And Practices For Inductive Coordination Of Electrical Power And Railroad Communication/Signal Systems, AAR Blue Book, 1977.

12-101

13
MITIGATION OPTIONS
This chapter provides comprehensive listings of mitigation options, including applicability, relative cost, and effectiveness. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE. Mr. Frazier has over 30 years of experience in the management and conduct of applied research programs and the application of the research results to the solution of system-specific problems. His major fields of experience include electromagnetic effects, electromagnetic compatibility, cathodic-protection systems, and detection systems. He is the author of the EPRI computer program CORRIDOR that has been extensively used by the engineering community to predict power-line interference on co-located structures. He is nationally known for his work on the electromagnetic compatibility of electric power, pipeline, and railroad systems. As President of Corr Comp Co., Mr. Frazier provides services to the power/pipeline/railroad industries in the pursuit of compatible common utility corridors. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American 13-1

Mitigation Options

Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA.

13-2

Mitigation Options

Introduction
Mitigation is what you do if you have, or suspect that you someday will have, a problem. Mitigation is something you change or add to allow the continued use of required systems and equipment. In cases of electromagnetic compatibility (EMC), mitigation is what you do to disrupt the source-path-receptor combination that allows undesired energy to cause problems. To successfully apply mitigation, the source, path, and receptor must first be identified. The reason for this is that different mitigation practices are only effective for certain situations. For example, some mitigation can be effective for transmission sources, but do nothing against distribution sources. Some mitigation is effective for magnetic induction, but ineffective against electric induction. This chapter provides a comprehensive list of mitigation measures. Each mitigation option is briefly described. Perhaps the most useful tool provided are the tables that show the relative effectiveness of each mitigation in various situations. A section is also provided on page 13-28 covering safe working clearances from energized electrical equipment. While this is not strictly mitigation of electromagnetic interference, it is highly relevant when mitigating potential hazards from electrical energy.

Mitigation Tables
Table 13-1 lists mitigation effective for cases of excessive common mode voltage. Table 13-2 lists mitigation effective for cases of equipment damage from power system events. Table 13-3 lists mitigation effective for cases of steady-state effects to equipment operation.

13-3

Mitigation Options

Table 13-1 Excessive Common Mode Voltage Mitigation Power System And Railroad Power System Mitigation Counterpoise Underground Counterpoise Aerial Load Current Limiting Phase Arrangement Optimization (Cancellation) Split-Phasing Phase Current Balance Optimization Phase Spacing reduction (Phase Compaction) Vertical Spacing Increase (Higher structures) Horizontal Spacing Increase Railroad Mitigation Reducing Length Of Electrically Continuous Track (At Power Frequencies) (Shorten Track Blocks) Counterpoise Underground Grounding Un-Signaled Track/Sidings (multigrounded rail can act as a counterpoise) Insulated Joint Inspection - Increased Insulated Joint Replacement - Shorted Or Leaky Units Surge Arresters Inspection - Increased Surge Arresters Replacement Failed Units By-Pass Shunts And Couplers Removal Resistive Network Installation Isolation Transformer Installation Working Gloved (Work using electrically insulated rated gloves.) Pole-Line Communication Systems Removed AC Shunts To Drain Interference To Ground (Not recommended - only use with extreme caution can cause track circuit unbalance.) 13-16 13-16 13-17 13-18 13-18 13-19 13-20 13-21 13-21 13-22 13-22 13-27 13-30 Page 13-7 13-8 13-9 13-9 13-10 13-11 13-12 13-12 13-13

13-4

Mitigation Options

Table 13-2 Equipment Damage from Power System Events Mitigation Power System and Railroad Power System Mitigation Counterpoise Underground Counterpoise Aerial Fault Current Limiting Neutral Wire Size Increase Vertical Spacing Increase (Higher structures) Horizontal Spacing Increase Automatic Reclosing Restriction Manual Reclosing Restriction Fault Current Clearing Time Reduction Railroad Mitigation Reducing Length Of Electrically Continuous Track (At Power Frequencies) (Shorten Track Blocks) Counterpoise Underground Grounding Un-Signaled Track/Sidings (multigrounded rail can act as a counterpoise) Insulated Joint Inspection - Increased Insulated Joint Replacement - Shorted Or Leaky Units Surge Arresters Inspection - Increased Surge Arresters Replacement Failed Units By-Pass Shunts And Couplers Removal Resistive Network Installation Isolation Transformer Installation Working Gloved (Work using electrically insulated rated gloves.) Track Circuit Lead Fusing Pole-Line Communication Systems Removed Ground Mats/Grids Around Equipment 13-16 13-16 13-17 13-18 13-18 13-19 13-20 13-21 13-21 13-22 13-22 13-24 13-27 13-30 Page 13-7 13-8 13-8 13-10 13-12 13-13 13-13 13-14 13-14

13-5

Mitigation Options

Table 13-3 Steady-State Effects to Equipment Operation Mitigation Power System and Railroad Power System Mitigation Counterpoise Underground Counterpoise Aerial Load Current Limiting Phase Arrangement Optimization (Cancellation) Split-Phasing Neutral Wire Size Increase Phase Current Balance Optimization Phase Spacing reduction (Phase Compaction) Vertical Spacing Increase (Higher structures) Horizontal Spacing Increase Open Delta Transformer Removal Capacitor Bank Inspection/Monitoring Railroad Mitigation Reducing Length Of Electrically Continuous Track (At Power Frequencies) (Shorten Track Blocks) Counterpoise Underground Grounding Un-Signaled Track/Sidings (multigrounded rail can act as a counterpoise) Insulated Joint Inspection - Increased Insulated Joint Replacement - Shorted Or Leaky Units Surge Arresters Inspection - Increased Surge Arresters Replacement Failed Units By-Pass Shunts And Couplers Removal Resistive Network Installation Isolation Transformer Installation Track Circuit Balance Optimization Battery Chargers Supplied at 240 Volts Operating Frequencies Chosen to Maximize Immunity Signal Output Level Increases Equipment Replaced with Less Susceptible Equivalent Systems Style C Track Circuits Removed 60 Hz Cab Signals Changed to Another Frequency Railroad Bed (Ballast) Condition Improvement AC Shunts To Drain Interference To Ground (Not recommended - only use with extreme caution can cause track circuit unbalance.) 13-16 13-16 13-17 13-18 13-18 13-19 13-20 13-21 13-21 13-22 13-23 13-23 13-25 13-26 13-27 13-28 13-29 13-29 13-30 Page 13-7 13-8 13-9 13-9 13-10 13-10 13-11 13-12 13-12 13-13 13-15 13-15

13-6

Mitigation Options

Detailed Options
Power System Mitigation Counterpoise Underground Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

In this context, a counterpoise wire is a multigrounded conductor used to reduce magnetic field levels through passive cancellation. (Power engineers usually think of counterpoise as a grounding method used in areas of low conductivity soil. This application is different.) When a changing magnetic field passes through a conductive loop (the counterpoise wire with the earth as a return path), current is magnetically induced in the conductor. That current creates a magnetic field of its own. That magnetic field opposes the field that originally created the current. The result is that net magnetic field levels near the counterpoise are lower than would exist without the counterpoise. A counterpoise wire in the ground can have the effect of disrupting the coupling path, thereby reducing the voltage magnetically induced in the rails. A counterpoise can also provide some protection against earth conduction by providing an alternate low impedance path to reduce current on the rails. The position of the counterpoise is critical to its function. Computer modeling is used to determine the optimum location, size, and grounding method for the wire. Care must be taken to insure that the counterpoise does not increase rail-to-rail voltages. The analysis must include not only the alignment of the wire, but how much variation in that alignment can be allowed during installation. Counterpoise wires are usually installed near the track, between the power line and the track. They can also be installed between tracks and on the side of the track away from the power line. Counterpoises usually require no maintenance. If designed with the proper materials, and correctly installed, corrosion should not be a problem. A break in a continuously grounded or multigrounded counterpoise only reduces its effectiveness by the percentage of the wire no longer carrying current. The usual cause of damage to underground counterpoises is dig-ins from work near the tracks. Any severed counterpoise wire shout be treated as an energized wire (shock hazard), and both ends should be grounded while repairs are completed.

13-7

Mitigation Options

Counterpoise Aerial Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

In this context, a counterpoise wire is a multigrounded conductor used to reduce magnetic field levels through passive cancellation. (Power engineers usually think of counterpoise as a grounding method used in areas of low conductivity soil. This application is different.) When a changing magnetic field passes through a conductive loop (the counterpoise wire with the earth and shield wires as return paths), current is magnetically induced in the conductor. That current creates a magnetic field of its own. That magnetic field opposes the field that originally created the current. The result is that net magnetic field levels near the counterpoise are lower than would exist without the counterpoise. A counterpoise wire on the poles can have the effect of disrupting the coupling path, thereby reducing the voltage magnetically induced in the rails. A counterpoise can also provide some protection against earth conduction by providing an alternate low impedance path to reduce current on the rails. The aerial counterpoise is usually mounted on the power poles below the phase conductors. Computer modeling is used to determine the optimum location and size for the wire. Aerial counterpoises have the effect or canceling some of the magnetic field near the source. Counterpoises usually require little maintenance. Visual inspection is typical. Electrical connections can be tested for high impedance. Any severed counterpoise wire shout be treated as an energized wire (shock hazard), and both ends should be grounded while repairs are completed. Aerial counterpoise should be marked as such so power company workers know that it is neither a phase conductor nor a neutral. Fault Current Limiting Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

13-8

Mitigation Options

Restricting or limiting the fault current on a power line has the effect of directly reducing magnetically induced energy in railroad circuit that parallel the line. Underground power cables often have large inductors in the substation to limit fault current. This is one of many reasons why underground power lines are much more costly than overhead lines. But, the cost of current limiting a power line is almost always prohibitive. Future developments for power delivery currently being studied may change this economic situation. If solid-state systems become commonplace in the power system of the future, fault current limiting may become more available. Load Current Limiting Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Load current limiting involves placing operational constraints on the power company to not exceed steady-state load levels that are known to be acceptable in a power line. This is more often a short-term measure used until permanent mitigation can be completed. However, if the power company does not need the full design capacity of the line, this measure could be permanent. The exact terms of load current limiting can be negotiated between the companies. Depending on the problems that result from higher loading, limits could be fixed or variable. Examples of variations could include emergency conditions on the power system, time of day variations, or seasonal variations. Phase Arrangement Optimization (Cancellation) Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Fault/Switching Steady State Common Mode Differential Mode Moderate Severe Extreme

On rights-of-way with multiple power circuits the phases of the circuits can be arranged so that steady-state magnetic fields at the ground are minimized. This is sometimes referred to as opposed phasing. Most power companies employ this practice as a matter of policy. This 13-9

Mitigation Options

practice reduces magnetic fields (a.k.a. EMF) as well as magnetic induction. It is fairly common for the fields from a double circuit line to drop to one third of their prior level through re-phasing. There are some older lines that are not phased to minimize ground level magnetic fields. One important consideration is that a line might be taken out of service for maintenance or through a fault. In such cases, the entire load might be carried on the remaining line(s) until normal operation is restored. The original higher levels of magnetic fields and magnetic induction would exist during this time. It is necessary to account for the effects of such a condition in evaluating the joint corridor. Also, some power lines are bi-directional, just as some tracks are bi-directional. If power lines are designed for maximum cancellation with current flowing one way, and one circuit switches direction of current flow, magnetic fields that were canceling may now add. Split-Phasing Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Split phasing is a way of achieving phase arrangement optimization (see above) on a single circuit power line. The line is constructed as a double circuit line. Each of the three phases of the power line is connected to one wire in each circuit. The phases are arranged for maximum magnetic field cancellation at ground level. Because the six phase conductors are all one line, the possibility of increased field levels with one side out of service is eliminated. One difficulty in implementing split-phasing is that the protective relaying that trips the line if there is a fault needs substantial modification to account for the multiple electrical paths to the fault (one direct down the faulted wire, one down the other wire of the faulted phase to the far end and back up the faulted wire). Neutral Wire Size Increase Application Sources: Paths: Problem: State: Mode: Cost: 13-10 Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Mitigation Options

When the source is distribution and the path is earth conduction from the multigrounded neutral, increasing the size of the neutral wire can reduce interference levels. The larger neutral wire has lower impedance. The neutral current is divided between the neutral and the earth in returning to the substation transformer. By lowering the impedance of the neutral, a larger portion of the neutral current will flow in the wire, and less will flow in the earth. The result is that the coupling path has been altered to reduce the ac interference in the rails. Phase Current Balance Optimization Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

The effect of unbalance is zero-sequence current. In distribution this is the current in the neutral wire (if there is one) and in the earth. This neutral/earth current can be a significant problem when the interference path is earth conduction. In transmission the magnetic fields from zero sequence current drop off proportionate to the distance. Magnetic fields from balanced currents drop off with the square of the distance. Additionally, some mitigation techniques do not work on zero-sequence current. Phase arrangement, split phasing, and phase spacing reduction are completely ineffective against zerosequence current. And, increasing the distance between the power line and the railroad is far less effective against zero-sequence current than balanced current (positive sequence). Every power company tries to balance phase current at the substation transformer. They do this because unbalanced phase currents reduce the capacity of the transformer, and can reduce the life of the transformer. But, for phase current balance to be effective in reducing magnetic induction the phase currents must be balanced everywhere along the line. Even if the power company did everything possible to balance every segment of a distribution line, there are three reasons why this effort can only have limited short-term success and little long-term success. First, single-phase loads tap into the line at discrete locations. Even assuming fixed loads at each customer, only so much can be done with scattered loads at various levels. Additionally, distribution lines usually split up when they are far from the source, with single-phase or two-phase branches heading farther out. These cannot be balanced. Second, new loads are frequently added to the power system. Often these are small loads that can be connected to the various phases so that the long-term effect of this growth does little to disturb the overall balance of the line. But, sometimes a relatively large load is added to a single phase (such as a subdivision with 100 homes). It may be impossible to offset the additional load by moving loads from phase to phase. The best we can hope for is to get as close as possible for most of the line. 13-11

Mitigation Options

Third, even if the line is perfectly balanced at the end of the day, and there are no major load additions, the moment-to-moment changes in customer loads will put the line out of balance. When home air conditioners come on the balance will change. When commercial buildings fill with workers and they all turn on their computers the balance will change. When the wind picks up and a load becomes a source because of a windmill the balance will change. It is the power customer, individually and as a group, that determines the amount of phase current unbalance at any given moment. In short, the power company should balance their lines as well as possible not just at substations, but all along the lines. However, those who are affected by unbalanced power lines cannot rely on phase current balance to solve their problems. Phase Spacing reduction (Phase Compaction) Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Reducing the spacing between phase conductors increases the effectiveness of magnetic field cancellation for balanced phase currents. This can be effective in reducing steady-state common mode magnetic induction. Cost on a proposed line might be moderate, while cost on an existing line could be extreme. Vertical Spacing Increase (Higher Structures) Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Increasing distance between transmission lines and track is usually an effective way to reduce ac interference, but it is often impractical. Increased vertical spacing requires taller structures. Unfortunately, many communities push for lower structures for aesthetic reasons. Sometimes phase compaction can be used on the same height structures to effectively increase the spacing to the bottom conductors. The effect of increased vertical spacing is reduced for the unbalanced current component on the line.

13-12

Mitigation Options

Horizontal Spacing Increase Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Increasing distance between transmission lines and track is usually an effective way to reduce ac interference, but it is often impractical. Increased horizontal spacing requires additional rightof-way. Blow-out easements1 might also be needed. Sometimes the wires can be strung on the far side of the structure to give some additional horizontal spacing (this can be a problem depending on the width of the right-of-way). The effect of increased horizontal spacing is reduced for the unbalanced current component on the line. Automatic Reclosing Restriction Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

A transmission line with high-speed protective relaying to clear faults usually automatically recloses the line after a very brief interval. If the fault was an arc, the ionized path will often disburse during the short time the line is off. Upon reclosing, the fault would be gone and the line is restored to service. Two or three reclosing attempts are typical. The problem occurs when the fault is physical something that wont just dissipate. Fault current will flow in the faulted phase, and into the earth at the point of the fault, with each reclosing. If the fault clears in 5 cycles and has two reclosings (each clearing in 5 cycles), then the fault current is present for 15 cycles or 0.25 seconds. Personnel safety from exposure to high voltages is related to the duration of the exposure. Longer exposure increases risk (see Chapter 7). Damage to equipment (such as surge arresters) 2 is dependent of the energy (I t, where I is current and t is time)(see Chapter 6). For both situations the total time for initial clearing and reclosing attempts is an appropriate value to apply. Analysis might determine that reducing the time will significantly reduce the risk to

Blow-out is when the wind blows the wire to the side. If the wire can cross the edge of the right-of-way such that is over private property, then additional easements might be needed.

13-13

Mitigation Options

both personnel and equipment. If that is the case, restricting the number of reclosings might be effective. Restricting the number of reclosing attempts will reduce the reliability of the transmission line. At least one reclosing attempt is usually needed to meet minimum reliability criteria for any line. If the initial fault destroys the arrester, then the subsequent fault current upon reclosing into the fault could damage signal equipment past the absent arrester or present a hazard to personnel. Manual Reclosing Restriction Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

When a transmission line locks out after automatic reclosing attempts fail, an operator will sometimes command a manual reclosing after a few minutes. If the fault was an arc, the ionized path will have finally disbursed during the extended time the line is off. If the fault still exists the line will trip off again and lock out. This additional fault time does not add to the other times as with automatic reclosing (the equipment has had time to cool and the personnel time to recover). The concern in this case is if railroad personnel are present and may have already begun repairs. If the arresters are no longer functional, an increased hazard could exist. Under such conditions, railroad personnel should consider working gloved (see Working Gloved later in this chapter). Fault Current Clearing Time Reduction Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Most transmission lines have high-speed relaying that leaves little room for improvement. So, this mostly applies to distribution. Distribution lines are often protected by fuses and reclosers in such a way that significant fault current can flow for several seconds or even continuously. It is rare, but not unheard of, for magnetic induction or earth conduction from distribution faults to damage equipment. New technology for distribution automation is being installed in many areas as a means to improve power system reliability. Some of these same technologies can be used to reduce fault clearing times on distribution lines. 13-14

Mitigation Options

Open Delta Transformer Removal Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Open-delta transformers2 are a means to provide three-phase service using two phases and two transformers. Properly applied open-delta transformers are not a problem. However, when the load applied to the open-delta exceeds design levels, one transformer saturates and substantial harmonic currents can be generated. Power companies use open-deltas in rural areas where customers want three-phase power, but are not willing to pay to bring all three phases of the distribution line to their location. Opendeltas are also applied for smaller three-phase loads because the installation is substantially less expensive than a larger three transformer set. In many areas electric companies are required by regulators to use open-deltas for smaller customers through their mandate to be as cost effective as possible. Unfortunately, many customers do a poor job of estimating their peak load. And, traditional electric meters give only average usage, not peak load. As a result, open-delta transformers are often overloaded and can become a source of conducted harmonic interference. Capacitor Bank Inspection/Monitoring Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Distribution capacitor banks that are not balanced can create harmonic currents that can couple into railroads through earth conduction. Usually capacitor banks become unbalanced because of an open fused disconnect on one phase or a failed capacitor. The first symptom of a capacitor bank problem is often a nearby railroad highway grade crossing false activation.

Open-delta transformer banks are also sometimes called open-wye transformers. In fact, they are usually open-wye on the primary side and open-delta on the secondary side.

13-15

Mitigation Options

Railroad Mitigation Reducing Length of Electrically Continuous Track (At Power Frequencies) (Shorten Track Blocks) Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

The amount of voltage induced into railroad rails with respect to remote earth depends heavily on the length of the exposure. The longer the track circuit - the greater the common-mode voltage that can be developed on the track. By limiting or reducing the length of electrically-continuous sections of track, we can reduce the induced voltage on the track. The length of electrically-continuous sections of track is determined primarily by the design of the signaling system and/or the propulsion return circuits (in the case of electrified railroads). Longer sections minimize the number of signaling circuits required, but also reduce the operating efficiency of the railroad by increasing the amount of time that each train spends in each block. As no other train may enter a block of track that is already occupied, longer blocks mean longer waits for other trains. The length of an electrically-continuous section of track may be reduced by adding sets of insulated joints to partition a section of track into a greater number of smaller pieces, or by removing bypass couplers used to carry signaling frequencies around insulated joints. Often this requires the installation of additional track circuit hardware. In the long term, partitioning the tracks covered by railroad signaling circuits into smaller pieces can contribute to higher maintenance costs and increased signal system complexity, but these may be slightly offset by increases in operating efficiencies and greater individual track circuit reliability (shorter track circuits usually have fewer problems with changing ballast resistance conditions). Counterpoise Underground Application Sources: Paths: Problem: State: Mode: Cost: 13-16 Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Mitigation Options

In this context, a counterpoise wire is a multigrounded conductor used to reduce magnetic field levels through passive cancellation. (Power engineers usually think of counterpoise as a grounding method used in areas of low conductivity soil. This application is different.) When a changing magnetic field passes through a conductive loop (the counterpoise wire with the earth as a return path), current is magnetically induced in the conductor. That current creates a magnetic field of its own. That magnetic field opposes the field that originally created the current. The result is that net magnetic field levels near the counterpoise are lower than would exist without the counterpoise. A counterpoise wire in the ground can have the effect of disrupting the coupling path, thereby reducing the voltage magnetically induced in the rails. A counterpoise can also provide some protection against earth conduction by providing an alternate low impedance path to reduce current on the rails. The position of the counterpoise is critical to its function. Computer modeling is used to determine the optimum location, size, and grounding method for the wire. Care must be taken to insure that the counterpoise does not increase rail-to-rail voltages. The analysis must include not only the alignment of the wire, but how much variation in that alignment can be allowed during installation. Counterpoise wires are usually installed near the track, between the power line and the track. They can also be installed between tracks and on the side of the track away from the power line. Counterpoises usually require no maintenance. If designed with the proper materials, and correctly installed, corrosion should not be a problem. A break in a continuously grounded or multigrounded counterpoise only reduces its effectiveness by the percentage of the wire no longer carrying current. The usual cause of damage to underground counterpoises is dig-ins from work near the tracks. Any severed counterpoise wire shout be treated as an energized wire (shock hazard), and both ends should be grounded while repairs are completed. Grounding Un-Signaled Track/Sidings (Multigrounded Rail can Act as a Counterpoise) Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

This mitigation technique is rarely used, but can be effective if there are already properly located dark (un-signaled) tracks adjacent to the affected track. This technique works much like the counterpoise described above, using the multi-grounded rails as the conductors of the counterpoise.

13-17

Mitigation Options

By grounding the un-signaled tracks, an opposing magnetic field is created around the un-signaled tracks by the induced current flowing through them. This opposing field can reduce the strength of the magnetic field around the affected track. Insulated Joint Inspection Increased Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Good railroad signal maintenance practices go beyond the minimums dictated by governmental regulatory agencies. Several railroads have their own local, regional, or system-wide standards for the testing of insulated joints. The effect of even marginally leaky insulation in an insulated rail joint can be significant, particularly in high ballast resistance conditions. Replacing or repairing defective insulation in insulated rail joints can significantly reduce the degree of electrical unbalance between the rails of a track. This in turn reduces the amount of differential voltage potential that can be created between the rails of a section of track. Due to the design of railroad signaling circuits, most track circuits will not experience operational problems until more than one insulated joint has either failed or become unusually leaky. But this should not be construed as grounds for ignoring individual insulated joint failures until they combine with other insulated joint failures to affect the operation of the signaling equipment. Increasing the frequency and effectiveness of insulated joint inspections and tests should be a priority for every railroad concerned about induced ac interference. And setting standards for insulated joint testing and remediation should be a part of every railroad signal departments maintenance plan. Insulated Joint Replacement-Shorted or Leaky Units Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

13-18

Mitigation Options

Shorted or leaky insulated joints contribute to rail voltage imbalance in ac induction environments by causing one rail of a track to have a rail-to-ground impedance that is distributed differently than that of the opposite rail. By replacing or repairing defective insulated joints, the rail-to-ground impedance of the two rails of a section of railroad track can be made more similar, resulting in less induced rail-to-rail (differential) voltage. It can also minimize the voltage potentials across other nearby insulated joints. Once an insulated joint has been determined to be defective, i.e. its electrical insulation is no longer effectively preventing the flow of electrical current from one rail to the other; it should either be replaced or repaired. Which approach is used will be determined by railroad policy, and by the type of insulated joint in failure. For example, glued insulated joints will often have to be replaced if they are found to be defective. In contrast, insulated joints constructed using urethane-encapsulated joint bars can be disassembled and checked for damaged or worn components. In some cases, the cause of the defect is as simple as wear products (metal particles) from the rail bridging the gap between the ends of the two rails to be electrically isolated. Cleaning or re-slotting the insulated joint may be all that is necessary. Surge Arresters Inspection Increased Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Periodic inspections of railroad signaling equipment installations should always include at least a visual inspection of the lightning arresters and equalizers on the surge panel of the installation. Any signs of significant arcing or damage to these devices should be grounds for further investigation. Additionally, the electrical characteristics of the arresters can often be quickly checked using a clamp-on ac current meter around the ground lead of the surge panel. Unless there is a shorted arrester on the panel, and an inductive electromagnetic field present at the rails, the current in the ground lead of the surge panel should always be essentially zero. The flow of as little as a few milliamperes of ac current may indicate the presence of leaky arrester which may fail in a much more substantial way in the future.

13-19

Mitigation Options

By replacing individual railroad surge protection devices before they create a significant rail-to-rail voltage potential by upsetting the electrical balance of the track, the potential for ac interference with the railroad signal equipment is minimized. Surge Arresters Replacement Failed Units Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Failed railroad surge protection devices (lightning arresters) can fail in one of two ways: either they become conductive when they are supposed to be in their non-conducting state to an unacceptable degree, or they fail to become conductive when they are supposed to be. The former is usually caused by the deposition of a layer of conductive material across the gap of the arrester, while the latter can be caused by the electrodes of the arrester being physically burned away to the point that the firing voltage becomes unacceptably high. In the extreme cases, the arrester either becomes shorted, or completely burned away (always open). A shorted arrester is one of the most common sources of track unbalance. There is no governmental requirement in the United States that dictates periodic testing of the surge protective devices on a railroad. Thus, shorted arresters may escape notice until the right inductive conditions occur to create a differential voltage (rail-to-rail potential) on the rails that is large enough to cause operational difficulties for the signal equipment. As shorted lightning arresters connected to the track can also cause large common-mode voltages to appear (rail-to-ground potential), this can also result in personnel safety concerns. The simple answer to a failed railroad surge protective device, whether it has become unacceptably conductive, or has been burned away, is to replace it. Most of the devices used have a parts cost of less than $50 (U.S.), and can be quickly and easily replaced. Some manufacturers recommend that any surge protective device connected either from rail-to-rail, or from rail-to-ground be replaced if it exhibits significant visual evidence of damage, or shows a resistance of less than 1000 Ohms. It should be noted that in the case of multiple failed surge protective devices, replacing a single device may temporarily appear to unbalance the track, until the remaining failed devices are found and replaced.

13-20

Mitigation Options

By-Pass Shunts And Couplers Removal Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

The purpose of removing by-pass shunts and couplers used to carry either a single frequency or a broad spectrum of railroad signaling frequencies around one or more insulated joints is to reduce the electrically-continuous length of the section of track in question. By reducing the length of the electrically-continuous section of track, its capacity to develop induced voltages will be reduced. Removing such by-pass shunts and couplers will usually necessitate the installation of additional signaling equipment often at significant expense to the railroad ($10K-$100K U.S. or more per location). For signal systems, this may require the installation of a cut section, where insulated joints are installed, and the section of track is partitioned into two separate lengths of track. For crossing warning systems, this may require the installation of additional Crossing Predictors or Motion Sensors, along with additional underground or overhead cable circuits. The long-term impact of such a change is increased maintenance cost for the railroad, but with proper design, the operational impact of this type of change can be made quite minimal. Resistive Network Installation Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Although rarely applied, resistive networks have been used to reduce high levels of both common mode and differential mode interference (see case study #1). Also see AC Shunts To Drain Interference To Ground at the end of this chapter.

13-21

Mitigation Options

Isolation Transformer Installation Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

In some cases, it may be desirable to isolate the railroads local ground from the power companys multi-grounded neutral. This can be done by providing an isolation transformer in the utility service drop to the railroad equipment enclosure from the utility distribution line. The result of this isolation transformer is to create a less-direct connection between the railroad ground, and the utilitys neutral wire a result of the earth resistance between the ground at the utility pole, and the railroads local ground. Thus, utility events which raise the potential of the neutral wire will have less effect on the electrical potential of the ground bus in the railroad equipment enclosure. This can help prevent railroad lightning arresters from firing backwards from the ground bus to the rails during a utility event in which the neutral-to-earth potential rises. Work Gloved (Work Using Electrically Insulated Rated Gloves.) Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Although it is highly unusual, some railroad workers have resorted to using insulated gloves when working on railroad signal circuits that suffer from extreme induced potentials. The effect of using this personal protective equipment is to insulate the worker from a voltage potential which may either be uncomfortable or unsafe to come in direct contact with. Although gloves are relatively inexpensive, the degree of safety that they provide is a function of both their consistency of use, and the magnitude of the electrical potentials being guarded against. Also, proper care and testing of rated electrically insulated gloves is essential to continued safety.

13-22

Mitigation Options

Track Circuit Balance Optimization Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

In general, most railroad track circuits are intentionally designed to be balanced. In order for a significant degree of unbalance to occur, something unusual must be at work. Electrically balancing the impedance of the rails of a track usually consists of repairing or removing the problem that is causing the unbalance. Removing an unbalance lessens the effects of induced voltages, because the induced voltages will tend to affect each rail equally. An impedance unbalance (with respect to earth) between the two rails of a track is the primary mechanism by which common-mode induced voltages are converted to differential voltages. As almost all railroad signal equipment is connected to the tracks in a differential fashion, reducing the differential voltage which results from induced common-mode voltages can improve the operation and reliability of the railroad equipment. This is a technique that applies equally to the design, construction, and maintenance of railroad signal circuits. Battery Chargers Supplied at 240 Volts Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

One often unrecognized potential source of interference on railroad track circuits is the power supply, battery charger, or rectifier used to both power the signal equipment, and keep the back-up batteries charged. Most of these battery chargers can be supplied from either 120 V ac or 240 V ac power, with the choice being made by wiring or jumper changes at the input terminals. A failed diode in the power supply will create ac harmonics that will flow back to the power company through the ac supply wiring. If such a charger is fed at 120 V on a system with a multi-grounded neutral then some of those harmonics will be conducted directly into the ground system of the railroad. 13-23

Mitigation Options

By switching from a 120 Volt to a 240 Volt supply, the harmonic energy will return to the power company on the phase wires not on the neutral. This isolation prevents the interference from coupling into the railroad systems. Track Circuit Lead Fusing Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

At present, fuses are rarely used in track circuit leads. However, their use is slowly becoming more common, particularly in areas which either have a history of signal equipment damage resulting from induced voltages on the tracks, or where electrified trains are run. The effect of fusing the track leads is to minimize the amount of energy that can be applied to the signal equipment from the tracks. Provided that the fuse acts quickly enough, and can interrupt the voltage present, fuses can significantly reduce signal equipment repair and replacement costs. In order to be effective, the fuse used must be carefully selected. The fuse must be capable of carrying the normal currents applied to or extracted from the track, and yet open quickly under conditions of abnormal current flows large enough to damage the signal equipment or pose a significant safety hazard to personnel. However, the risks associated with the step-and-touch voltage potentials created may not be easily evaluated, and few equipment manufacturers provide specifications for their equipment indicating the amount of energy (voltage and/or current) they can withstand for brief periods. That being said, one common approach to the fusing of track leads in electrified-propulsion territory has been to install cartridge-type fuses rated for 600 Volts and and a current interruption rating of 10,000 Amperes or so. These units are often of the rebuildable type, with a low-cost replaceable element installed inside of the cylindrical body. These are just over 1/2-inch in diameter, by a few inches in length. The current rating of the fuse element is chosen to easily carry the intended signal circuit current, provide a low-enough resistance so as not to significantly affect the operation of the signal circuits, and still open quickly during an excessive-current event. Often this leads to the selection of a fuse with a rating of a few tens of amperes. The fuses are installed in series with the track leads, between the tracks and the surge panel of the railroad signal equipment installation. This way, the firing of an arrester or equalizer can serve as a trigger to open the fuse element.

13-24

Mitigation Options

One manufacturer of SPD panels has incorporated a small gage jumper wire that acts as a fuse link to protect the equipment. The downside to using fuses is that there may be some degree of nuisance opening, in which some form of surge event may open a fuse, leaving the railroad signal equipment in an inoperative but fail-safe condition. This may impede normal railroad or grade crossing operation, but is often preferable to damaged signal equipment that is expensive both in terms of time and money to repair or replace. Operating Frequencies Chosen to Maximize Immunity Application Sources: Paths: Problem: State: Mode: Cost: TransmissionDistribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment Damage Equipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

The careful selection of operating frequencies for railroad signal equipment is one of the easiest ways of avoiding ac interference problems. By using frequencies that are as far away as possible (in frequency) from the harmonic components of ac interference, the impact of the unwanted ac energy can be minimized. Most railroad signal equipment using ac audio frequencies is essentially narrowband in nature. That is, the electrical signals applied to the railroad tracks occupy only very small range of frequencies often spread over just a few Hertz about a particular center frequency. Thus, their receivers are usually equipped with very narrow filters that admit only the operating frequency of the desired transmitter. The receiver filters greatly attenuate all other frequencies being received from the track. However, the amount of attenuation that will be applied to any unwanted or unintended ac frequency on the track is never infinite, and some small amount of the undesired frequencies always gets through. This amount is increased if the interfering frequency is close (in frequency) to the operating frequency of the signal equipment. Since the frequency of the interfering ac energy can seldom be changed at will, it is often much easier to select operating frequencies for the railroad signal equipment which are as far away as possible from the interfering ac energy primarily the power-line frequency and its significant harmonics. Unfortunately we cannot choose railroad signal equipment operating frequencies at will. Due to the attenuation characteristics of railroad track, the higher the operating frequency, the shorter the maximum length of the track circuit. Often this frequency selection process seems paradoxical. For example, an operating frequency of roughly 156 Hz can be interfered with by the 180 Hz harmonic of a 60-Hz power system. Choosing an alternative frequency such as 114 Hz may at first appear to be a bad idea, since it 13-25

Mitigation Options

would seem to be even closer to the 120 Hz harmonic of 60 Hz. However, since the 120 Hz component of the unwanted ac energy on the tracks is often much smaller in amplitude than the 180 Hz component, this change can provide a significant improvement in interference immunity. The key to success is to always consider the frequency spectrum of the unwanted ac energy, and compare it to the operating frequencies and receiver bandwidths of the signal equipment. If at all possible, this should be done during the design stages of a project, in order to avoid problems from the outset, rather than having to re-allocate signal equipment frequencies after the signal equipment is already in operation. Signal Output Level Increases Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Most cases of ac interference can be described in terms of the signal-to-noise ratio, which compares the amplitude of the signal to be received, with the amplitude of all other signals present. The stronger the intended signal, with respect to the other noise signals, the higher the signal-to-noise ratio, usually expressed in decibels (dB). One way to improve the signal-to-noise ratio is to increase the amplitude of the electrical energy transmitted by the railroad signal equipment. Many different types of track circuits offer the option of using a higher transmit level (voltage, current, or power). This is normally compensated for by using a greater degree of attenuation at the receiver, in order to keep the received amplitude the same. This greater attenuation in the receiver not only keeps the amplitude of the desired electrical signal the same, it also applies a greater degree of attenuation to all of the other unwanted electrical signals coming from the track. Thus, the signal-to-noise ratio is increased. Improving the signal-to-noise ratio of the railroad signal equipment often results in a measurable improvement in railroad signaling equipment operations. However, this must be done with caution, as the increased transmit level used may also decrease the signal-to-noise ratio for other railroad signaling equipment. Signal equipment manufacturers guidelines must be carefully consulted before making such changes, as a seemingly simple change can have wide-reaching effects, and even signal system safety considerations.

13-26

Mitigation Options

Equipment Replaced with Less Susceptible Equivalent Systems Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

After exhausting more economic mitigation options, the one remaining option is to change the characteristics of the receptor. For each specific railroad signaling task performed, there are often several different types, and different brands of signaling equipment to choose from. Each has its own particular advantages and capabilities. There are generally two distinct motivations to change signaling equipment in cases of ac interference. First - due to the durability of railroad signaling equipment, much of the equipment currently in use is years or even decades old. Upgrading to newer equipment can allow railroads to take advantage of the significant improvements in electronic signal processing technology that have been made in the past few decades. Many newer designs have significantly greater immunity to ac interference. Replacing antiquated signal equipment can also bring added benefits to the railroad in the form of increased equipment reliability. The second motivation for upgrading signal equipment in ac interference environments is that making a fundamental change in the technology used to perform a particular signaling function can provide large improvements in ac interference immunity. For example, changing from Motion Sensors to Crossing Predictors can provide a significant increase in ac interference immunity. Because of differences in the way they process and interpret the electrical voltages on the tracks, Motion Sensors - and in particular older Motion Sensors are generally much more sensitive to the influences of ac power-line-related energy than Crossing Predictors. A careful examination of the functional requirements of the affected railroad signal equipment should form the basis for any changes or upgrades. In some cases, add-on filters or other upgrades and accessories may be all that is needed. Pole-Line Communication Systems Removed Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme 13-27

Mitigation Options

Due to their frequent physical proximity to nearby ac distribution and transmission lines, pole-line communications systems (the telegraph wires found along railroad tracks) are often affected by induced interference from ac power lines. However, such open line-wire circuits are rapidly being replaced by other alternatives such as coded track circuits, vital radio communication systems, and even fiber-optic communications. This is being done for a variety of reasons, with increased performance and reduced maintenance cost chief among them. By communicating track occupancy information along a railroad line by other methods that are more resistant to ac interference, the problem of induced interference on pole lines can be effectively eliminated. However, the cost of this type of signaling equipment change-out can be quite large. If present trends continue, it seems likely that very few pole-line-based railroad signaling communication systems will be in use within a few decades. In some cases, replacing these systems with alternative technologies on an accelerated schedule may provide a means for resolving some ac interference problems in a way that is acceptable to both the electric utility and the railroad. Style C Track Circuits Removed Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Style C track circuits can use power-line frequency electrical energy to detect the presence of trains within a defined section of track. They are often used in areas where poor shunting is a concern, due to their higher rail-to-rail voltage. A favorite application for them is their use in lightly-used grade crossing warning systems at rural grade crossings, due to their low cost and good shunting performance. However, since these circuits often use ac energy derived from the electric utility (applying it at a reduced voltage of 6 Volts ac rms or less), this creates the potential for both interference with and damage to railroad signal equipment as a result of an unusually direct connection between the railroad tracks and the electric utility. This is an example of railroad signal equipment creating an interference problem that may appear to be the result of ac induction, but is really a case of conducted interference and signal equipment incompatibility. The second problem with the Style C track circuit is the fact that although they do not require any electrically-powered equipment at one end of the circuit, they achieve this by placing a rectifier diode directly across the tracks. This diode performs a half-wave rectification of the ac source when the track circuit is unoccupied. This rectification results in a broad spectrum of source-frequency-related harmonics. Style-C track circuits are typically isolated from the rest 13-28

Mitigation Options

of the track by pairs of insulated joints at both ends of each circuit. But any other signaling equipment located within the boundaries of the Style-C track circuit may be affected by the harmonics generated. One option is to use a Style C track circuit that uses another frequency other than the ac utility frequency. In theory, this type of track circuit could be designed to use almost any reasonablylow audio frequency, provided that it would still support track circuits of sufficient length. One manufacturer offers Style-C track circuit equipment that operates at 156 Hz. Changing the operating frequency of the Style-C track circuit may eliminate the harmonic content interfering with other signal equipment. But there are many other track circuits that could be used to replace a Style-C track circuit. Some have additional installation requirements, such as requiring additional track wires, or equipment shelters at remote locations, but there are many alternatives to choose from. 60 Hz Cab Signals Changed to Another Frequency Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

With respect to ac interference, the use of 60 Hz energy for Cab Signaling systems has seemingly obvious problems. However, the economic advantages of a railroad signal circuit that has such a simple transmitter (a transformer, resistor, and coding relay) have led to a limited number of installations in the United States. As the receivers used in such a system are intentionally designed to be able to detect 60 Hz energy, they are particularly susceptible to interference from ac power lines. The obvious solution to this problem is to change the operating frequency of the cab signal equipment. Railroad Bed (Ballast) Condition Improvement Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme 13-29

Mitigation Options

One mechanism for turning induced common-mode voltages on railroad tracks into interfering differential-mode voltages is the unbalance caused by uneven ballast resistance. If the rail-toground resistance of one rail of the track is grossly different than the other, this will result in a conversion of induced common-mode voltages into differential-mode voltages. By performing proper ballast maintenance, the ballast resistance (as measured from each rail to ground) will tend to be more even, thus minimizing any voltage unbalance between the rails. Ground Mats/Grids Around Equipment Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

It is common practice to install a bare wire in the earth around signal equipment enclosures. This wire is bonded to ground rods in the corners and the enclosures primary ground. This simple system effectively reduces touch potentials a person could encounter standing next to the enclosure while touching its conductive surface. By providing a conductive path between the enclosure and the ground immediately around it, the touch potentials are reduced. This type of system could be considered for any conductive apparatus that has been shown to present a potential hazard. Additionally, a ground mat can be added to such systems at doors and access panels as necessary. AC Shunts to Drain Interference to Ground (Not recommended - only use with extreme caution can cause track circuit unbalance.) Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Attempts to drain ac interference to ground with ac shunts must be applied carefully to avoid creating track circuit unbalance. The shunts must be installed in matched pairs (one to each rail) to a common ground. If the two shunts are not identical in electrical characteristics, they may cause unbalance. This solution has often been known to cause more problems than it fixed.

13-30

Mitigation Options

Never install a single ac shunt from rail to ground in an effort to reduce ac interference. This creates a track circuit unbalance. Never use a single shunt to ground in an effort to balance ac interference in a track circuit. Instead, find the cause of the unbalance and repair/remove it. It is extremely unlikely that adding a rail-to-ground shunt to reduce unbalance will work over time and changing conditions. Safe Working Clearances from Power Lines Application Sources: Paths: Problem: State: Mode: Cost: Transmission Distribution Railroad Magnetic Induction Electric Induction Earth Conduction Personnel Safety Equipment DamageEquipment Operation Steady State Fault/Switching Common Mode Differential Mode Moderate Severe Extreme

Various standards exist establishing safe working clearances from power lines. In the United States OSHA sets such standards. Always check with the local power company for their working clearances for untrained workers (their standard clearances may be greater than OSHAs). Remember that the trained workers with reduced clearance requirements are power company workers with current training. This normally doesnt include railroad signal personnel. Figure 13-1 shows an example of a working clearance card distributed by an electric company. Figure 13-2 shows as example of a signal installation with marginal working clearance from the top platform. No one is sure how they managed to install this equipment.

Figure 13-1 Sample of a Working Clearances Card

13-31

Mitigation Options

Figure 13-2 Railroad Signal Equipment with Marginal Working Clearance

13-32

14
CASE STUDIES
Introduction
Case Studies are provided to help the user of the Handbook open his or her mind to the wide range of possible causes of problems, and mitigation available. As the user follows the investigative process from Chapter 9, using the diagnostic flow chart from Chapter 10, the problem will usually be narrowed down to one of the four interim conclusions: Excessive common mode voltage Equipment damage from power system events Steady-state effects to equipment operation Problem is not caused by ac interference

The Case Studies are grouped in this same way. Each case study begins with the list of questions (diamond shaped decision boxes) from the diagnostic flow chart. This should allow the user to focus in on the case studies that most closely follow the flow chart path of the problem at hand. In the process of assembling the case studies for this handbook, and formatting them to parallel the diagnostic flow chart, it was noted that virtually every case study included data that was unnecessary. In several cases the extra measurements constituted the majority of the information provided, and the time expended in studying the problem in the field. The lesson learned is that the diagnostic procedures presented in this handbook will save time and reduce costs.

14-1

Case Studies

Excessive Common Mode Voltage


Case Study #1 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? _Yes_ _____ _____ _____ _60Hz _Yes_ _____ _____

Problem: False activation of highway grade crossing equipment and excessive rail-to-ground voltage General Description of Problem: The Power Company obtained a license agreement from a railroad for a ten-mile segment of a future 345 kV double circuit steel pole line. The first phase of the project called for a single circuit of 1590 MCM ACSR bundled conductor, which was pulled in vertical configuration on the structure side nearest the rails. The line was parallel and fairly close horizontally to the rails. Symptoms: Shortly after the line was energized, the railroad crossing equipment began to malfunction. The crossing guards would operate even in the absence of trains. Observations: This situation was modeled using an early version of the EPRI Corridor program. It predicted the rail-to-rail voltages would be 3 volts. The measured rail-to-rail voltages along the railroad were close to the calculated values except at one location near a spur where the measured values were significantly higher Measurements: The voltage across some insulating joints was 115 volts. Measurements indicated the rail-to-ground voltage was 68 volts at certain locations. The transmission line was de-energized due to the voltages apparently induced in the rails. Diagnosis (source, path, & receptor): The rail-to-rail and rail-to-ground voltages needed to be reduced. Corridor was used to look at many methods to reduce rail voltages. Some of the more promising ones were:

14-2

Case Studies

1. Use of additional shield wires 10 below conductor 2. Buried conductor in the ground near the rails 3. Raise transmission conductor ten feet 4. Shorten rail track sections 5. Move transmission line ten feet further away from rails 6. Resistive network on end sections of rail track. 7. Add second circuit with different phase configuration ABC CBA top to bottom. Methods six and seven were evaluated as being the most practical for this situation. Mitigation (options and chosen mitigation): This railroad situation was modeled using an early version of the EPRI Corridor program. It predicted the rail-to-rail voltages would be 3 volts. The measured rail-to-rail voltages along the railroad were close to the calculated values except at one location near a spur where the measured values were significantly higher. The railroad replaced a shorted insulator joint at this location and the rail-to-rail voltages dropped significantly. Further mitigation of induced voltages was accomplished by the following: Add a Resistive Divider Network at the Ends of Rail Sections: The Corridor program indicated that the rail-to-ground voltages could be reduced significantly by the use of a resistive divider network of 0.5-ohm resistors to ground installed at each end section of track. Voltages should drop from 100 volts to ground to 25 volts to ground. The Corridor program indicated that rail-torail voltages should be lowered to values of 1.6 volts with the resistive network. The railroad installed the restive network. The voltage measurements were close to predicted values. Various tests were made to assure the safety and operation of railroad equipment at the time the resistors were installed. This was a quick fix as it got both the railroad equipment and transmission line back in service. Corridor predicted that future increased loading on the line would raise rail-to-ground voltages above 40 volts. A second transmission circuit was needed. Add Second Circuit: The Corridor Program was used to determine the proper phasing for the second circuit to minimize induced voltages on the rails. Corridor predicted the second circuit would lower voltages to 5 volts rail to ground and 0.6 volts rail-to-rail when used in conjunction with the resistors. The Power Company pulled in a parallel second circuit. The phases nearest the rails were ABC top-to-bottom and the second circuit was pulled in with CBA top-to-bottom phasing. This significantly reduced both the rail-to-rail voltages and the rail to ground voltages. The actual measured values were even lower than Corridor predicted.

14-3

Case Studies

Case Study #2 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? _Yes_ _____ _____ _____ _60Hz _Yes_ _____ _____

Executive Summary: Mitigation was installed for a new transmission line parallel. Yet, before the line was energized, the railroad experience false activations at crossing along the route. Part One The Proposed Line Problem: The power company purchased an easement for a pole mounted double circuit 138kV transmission line to parallel a single track railroad for seven miles. The railroad required the power company to mitigate any ac interference/safety issues. Details: The proposed line was a compact structure (short and with phases close together) to conform to local community requests. This placed the wires closer to the track, increasing magnetic inducting both steady state and under fault conditions. As a result, mitigation was particularly complex. The parallel was modeled using the consultants proprietary software, and many mitigation options were evaluated. The selected mitigation was also modeled in Corridor as a check of the consultants software. The mitigation included: Removing the open wire signal pole line Installing Electro Code track signaling Replacing relay based crossing systems with electronic systems (PMD-2) compatible with Electro Code Installing a buried signal cable Installing counterpoise wire underbuilt on the transmission poles Installing counterpoise wire underground on both sides of the track Installing counterpoise wire on wood poles near the source substation on the side of the tracks opposite the transmission line

The driver for the extensive mitigation was the high level of fault current magnetic induction.

14-4

Case Studies

After all the mitigation had been installed and the transmission line built but, before the line was energized the railroad notified the power company of chronic false activation problems at two crossings near the source substation. Part Two False Activations Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: False activations of PMD-2 crossing equipment due to ac interference. Executive Summary: Harmonic earth currents converging on the substation from the surrounding community caused the false activations. This interference existed prior to any work being done at the site. The old relay grade crossing systems were immune to the interference. But, the new audio frequency electronic systems were susceptible. Detail: The power system was inspected and tested throughout the area to see if any problem existed that could have resulted in elevated neutral earth current or elevated harmonics. No problems were found. The interference levels present were essentially normal for this type of suburban community. The PMD-2s were replaced with PMD-3s, and later with HXP-3s. Ultimately, the HXP-3 units withstood the harmonic rich environment with carefully selected operating frequencies. Lesson Learned: Whenever replacing old signal equipment with electronic systems, measure the rail-to-rail ac spectrum. Forward this data to the signal equipment manufacturer for recommendations of equipment and installation. __No_ __No_ _____ _____ _____ _____ __No_ _Yes_

14-5

Case Studies

Case Study #3 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: Railroad workers feel tingling from the rails. Executive Summary: Steady-state ac voltage on the rails of the single track line measured as high as 90 Vrms across an insulated joint in a 7 mile single circuit transmission parallel. Mitigation involved stringing additional conductors on the transmission poles. Details: The transmission line was strung on double circuit poles. The poles were designed to carry two circuits in vertical delta configuration. One circuit was on each side of the pole. On each circuit, two phases were on a long lower arm. The third phase was on a shorter upper arm. Because of the weight of the phase conductors, two phases were strung on one side (upper position and outside position on lower arm), and the third phase was strung on the other side (outside position on long arm). To string all three phase conductors on one side, with the other side vacant, would have exceeded the structural capability of the pole. The problem with this configuration was that the distance between phase conductors is over 60 feet. Magnetic field levels were significantly higher than a typical line, and magnetic induction was a problem. Because a second circuit was going to be strung in a few years, the most cost effective mitigation was to string the other three conductors for the length of the parallel corridor. The active phase wires would be connected so that all three would be on the same side of the pole (away from the track). The other three wires would be grounded at the ends to allow them to act as counterpoise wires, and further reduce magnetic induction. A rejected option was to use the extra three wires to create a split-phase configuration (two wires for each phase, arranged to maximize field cancellation). The problem with this often-effective technique was that the protective relaying on this line could not handle the two paths to any fault within the split-phase section. (If A-phase faults to ground, there is a direct path down the faulted wire, and there is the longer path down the other A-phase wire to the end of the phase split and back on the faulted wire.) Reworking the relay protection scheme was not a desirable option to the power company. 14-6 _Yes_ _____ _____ _____ _60Hz _Yes_ _____ _____

Case Studies

Equipment Damage from Power System Events


Case Study #4 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: Power transmission fault currents destroy track circuit surge protective devices (arresters). Executive summary: A series of transmission line faults were triggered by lightning, road debris and salt falling into the lines (they went under the highway), and equipment failures. Fault currents were unusually high because of nearby generating stations. Despite extensive mitigation, the SPDs were still considered sacrificial in the case of power line faults (i.e. they were expected to fail leaving an open circuit to ground). The railroad installed fuses on the track circuit leads. Details: The 10 mile right-of-way had from 2 to 4 mainline signaled tracks. The initial power line fault would blow out most of the arresters. Occasionally, reclosing into a fault would then damage equipment. (Automatic reclosing will restore the line if the fault was a transient event that has cleared. Reclosing into a fault will result in high fault currents and rapid tripping of the line, again.) On one occasion, the signal maintainer was replacing blown arresters from an event a few hours earlier, when the line tripped again blowing the arresters he had just installed. The maintainer was not injured. Automatic reclosing of faults on the transmission circuits paralleling the tracks was reduced from the normal two attempts to one. An order was placed on the power systems operators not to order a manual reclosing on the lines until an inspection had been made. While plans were made to relocate the transmission lines away from the highway, the railroad decided to install fuses on its track circuit leads. These fuses would open the track circuit under fault surge conditions. While track signals would be lost until the fuses were replaced, the signal equipment was effectively, protected. Even the SPDs were protected from damage. It should be noted that at least one manufacturer of surge protection panels incorporated a thin wire jumper between the track lead and the arrester. This jumper acts as a fused link, melting and opening the track circuit before the arrester is destroyed. 14-7 __No_ _Yes_ _Yes_ _____ _____ _____ _____ _____

Case Studies

Case Study #5 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: Distribution line fault currents destroy track signaling equipment. Executive summary: Storm induced faults in distribution lines caused damage to signaling equipment. Details: The railroad contacted the power company with a claim for replacement of signal electronics destroyed by induction. The two events happened about one week apart at two different locations (about 5 miles apart) along the same rail line. In each case arresters were blown at several locations with equipment damage a one location. Distribution lines paralleled the entire stretch of track. The lines were heavily loaded and had relatively high fault current potential. The power company reviewed its records of distribution line events and discovered that the line paralleling the track had been hit by lightning and faulted at times corresponding with each reported incident. Additionally, the locations of the power line problems corresponded with the railroad equipment damage. It is unclear if the damage to the railroad was caused by magnetic induction or conducted fault current, but the correlation to power company events was clear. It should be noted that while railroad equipment damage from distribution line events is rare, the distribution wires tend to be closer to the track increasing coupling, and the fault clearing times tend to be much longer than for transmission lines. Both these factors increase the energy affecting the equipment. No mitigation was implemented. Normal maintenance continued on both the power lines and the railroad. No further losses have been experienced along this right-of-way. __No_ _Yes_ _Yes_ _____ _____ _____ _____ _____

14-8

Case Studies

Case Study #6 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: Power transmission fault currents destroy track circuit surge protective devices (arresters). Executive Summary: After major electrical storms the signal maintainer often finds blown arresters. He thinks it is more common along a transmission line parallel. They have not lost any other equipment. Arrester sweeps after storms have become standard procedure. Details: In the great plains, electrical storms are common. Transmission towers are the tallest things out there, and they get hit regularly by lightning. Sometimes an insulator will flashover (or back flashover) and cause a fault. The line trips off and then recloses. In the case of lightning caused trips, the line will usually close in on the first attempt. This scenario leaves track circuit arresters (SPDs) blown, but track circuits operational. Some signal engineers refer to this as the sacrificial nature of arresters. __No_ _Yes_ _Yes_ _____ _____ _____ _____ _____

14-9

Case Studies

Case Study #7 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: During the morning rush hour, an insulated joint on a commuter rail line melted. Executive Summary: Lightning caused a substation fault. Primary high speed relaying had been taken off line for maintenance. The current induced in the rail caused an insulated joint to melt away. Power company operating rules were changed to reduce the chances of a reoccurrence. Details: A double circuit 138kV transmission line paralleled the double track commuter line for about 6 miles to a substation. The conductors of the closest circuit were directly above the outside rail. The communication system that is critical to the operation of high speed relaying had been having problems and maintenance was scheduled. On the appointed day the power company communications people took the communication line, and the corresponding high speed relaying, out of service for maintenance. This was done despite the severe thunderstorm in progress at their location. When lightning hit the transmission line, a lightning arrester on one of the substation transformers directed the energy into the substation ground grid. Normally, the fault would be cleared in less than 5 cycles (0.083 seconds). With the communication line disabled, the fault did not clear for 45 cycles (0.75 seconds) using back-up relaying. During that time the lightning arrester welded the phase conductor solidly to the ground grid. When the relay attempted to reclose the line into the fault, fault current flowed for another 30 cycles (0.50 seconds) until it tripped again and locked out. The phase that faulted was the bottom phase on the circuit closest to the tracks. For the duration of the fault (1.25 seconds total) the current induced in the rail, as estimated using the Corridor software program, was 6,000 Amperes. It took a trained eye to see that what was left used to be an insulated joint. Lessons Learned: Never perform relay maintenance in a thunderstorm. Additionally, the Corridor study indicated that induced voltages under fault conditions even with high speed relaying operational were higher than desired. It was decided to install a counterpoise wire on the transmission poles below the phase conductors. Additionally, automatic reclosing of faults on the transmission circuits paralleling the tracks was reduced from the normal two attempts to one. 14-10 __No_ _Yes_ _Yes_ _____ _____ _____ _____ _____

Case Studies

Steady-State Effects to Equipment Operation


Case Study #8 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: False activations of PMD-2 crossing equipment due to ac interference. Executive Summary: The team was able to identify that one rail was grounded. This resulted in unacceptably high third harmonic rail-to-rail levels. The ground was caused by an improperly seated motion detector surge arrester in the railroads signal bungalow. When the surge arrester was placed in its socket correctly, the rail-to-rail ac voltages dropped to insignificant levels, and the false activations were eliminated. Detail: The railroad employees had been working on this problem since it first appeared two months earlier. When a train approached from the East (compass South) and stopped within the crossing threshold, the gate should go up after about 30 seconds. The PMD-2 monitors the 210 Hz signal phase to determine that the train is stationary. When the problem was present the phase had a jitter of about 7 degrees of phase angle - sometimes triggering false activations. The railroad had measured ac rail-to-rail voltages as follows: 60 Hz +10 dB 120 Hz -35 dB 180 Hz -16 dB 210 Hz (signal) -14 dB __No_ __No_ _____ _____ _____ _____ _Yes_ _____

The third harmonic level is high enough to get through the receiver filter and cause the phase variation noted. Power company measurements were: 60 Hz ----120 Hz ----180 Hz 0.10 Vrms 210 Hz (signal) 300 Hz 0.14 Vrms 0.11 Vrms

These numbers confirm the railroad measurements.

14-11

Case Studies

Other measurements showed that one rail had almost no ac voltage present while the other rail had rail-to-ground voltage equal to the rail-to-rail voltage. This condition is not common because whatever causes voltage to be present on one rail usually causes similar voltage to be present on the other rail as well. The hypothesis was that one rail was grounded. That put us in the questionable position of trying to convince railroad personnel that we need more ac voltage on one of their rails in order to make it work correctly. We were able to trace the ground into the signal bungalow, past the front end surge protectors, past the wiring harnesses to the Motion Detector Surge Arrester panel. When the surge arrester was removed from the holder the rail ground was eliminated. The surge arrester proved to be good, but when it was reinserted the ground was still gone! We were then able to determine that the surge arrester had previously been improperly installed. (The signal maintainer initially removed the surge arrester with his fingers. It is impossible to remove the surge arrester with your fingers when it is properly installed. A tool such as a screwdriver is needed.) We noted that with the ground eliminated, the phase jitter was eliminated, and the false activations were eliminated. We then repositioned the surge arrester, as we believe it had previously been misaligned, and the jitter returned. A West bound train conveniently stopped just East of the crossing allowing us to reproduce the problem and correct it at will. With the PMD-2 phase readouts stabilized the rail-to-rail ac voltages were dramatically reduced, while the signal at 210 Hz remained unaffected. Case Study #9 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: False activation of highway grade crossing equipment General Description: The power company has a single circuit 138 kV wood pole transmission line (delta configuration) with distribution under build (flat configuration) that is 85 feet from the center line of a railroad section. The transmission line parallels the railroad for about 0.4 miles. __No_ __No_ _____ _____ _____ _____ __No_ _Yes_

14-12

Case Studies

The section of railroad is 1.58 miles long and has six highway signal arm crossings. This transmission line had been in service for many years. Symptoms: The Power Company received notification from the railroad that several crossing arms would operate without the presence of a train. This occurred periodically over a period of the past 3 years. The failure of crossing arms to operate properly would occur at various times day or night. Observations: This situation was modeled using EPRI Corridor program. A maximum expected load for the transmission line was used to calculate rail voltages. The calculated rail-torail voltage was 0.3 volts as compared to 0.5 volts at the end section of the railroad. At the same location the calculated volts rail-to-ground was 11 volts as compared to a measured value of 5 volts. Measurements: A signal specialist from the railroad measured values of insulation resistance joints and found them to be within their accepted limits. The measured rail-to-rail voltages of 0.5 volts were considered to be within acceptable operating limits for signal equipment. Diagnosis (source, path, & receptor): Since the voltages on the rails were within the acceptable limits of railroad equipment a close evaluation was made of the rail equipment, crossing signal frequency settings and rail shunts. Mitigation (options and chosen mitigation): As a result of the railroad specialists looking at the rail situation the following actions were taken: 1. All 60 & 180 hertz capacitive shunts were removed from various crossing locations on this section of track. (The effect of the shunts had been to short out some of the signals used by crossing equipment) 2. Obsolete gas tube arresters were removed from crossing equipment. 3. Two of the lower frequency crossing equipment were replaced with HXP-3 crossing equipment. 4. All of the signal crossing frequencies were re-aligned to higher frequencies. No further problems have been reported with railroad crossing signal systems on this section of the railroad.

14-13

Case Studies

Case Study #10 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: False activation of highway grade crossing equipment Executive Summary: Every evening for the past three days the gates have come down when no train was present. The problem always happens between 6 pm and 8 pm and goes away by 9 pm. A failed capacitor was the cause of the problem. Details: As a result of past experience, the power company has a practice of inspecting all the distribution capacitor banks near the affected site of false activation. Five capacitor banks were identified from system maps. Inspection of the capacitor banks revealed a failed capacitor (thermal camera identified it at ambient temperature, while the other capacitors were at about the same elevated temperature). The capacitor bank with the failed capacitor switched on a timer. It was in service from 7 pm until 7 am each day. A capacitor bank with a failed phase injects high levels of harmonic current into the system and into the earth. These harmonics caused the problem. A question remained as to why the problem stopped by 9 pm each day. A second capacitor bank was identified that switched based on line voltage. Voltage records showed that this capacitor bank had switched in between 8:30 and 9 pm during each of the prior three days. When this capacitor bank was in service it somehow reduced the harmonics enough to protect the railroad. Mitigation: The capacitor bank with the failed unit was taken out of service until the capacitor could be replaced. __No_ __No_ _____ _____ _____ _____ __No_ _Yes_

14-14

Case Studies

Figure 14-1 Capacitor Bank

Case Study #11 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: False activation of highway grade crossing equipment Executive Summary: The power company received a complaint from the railroad that crossing gates were down due to ac interference since the day before. (Apparently, it had taken 24 hours for the power company call center to figure out who should take care of the problem.) A blown fused disconnect on a capacitor bank was the cause. 14-15 __No_ __No_ _____ _____ _____ _____ __No_ _Yes_

Case Studies

Details: As a result of past experience, the power company has a practice of inspecting all the distribution capacitor banks near the affected site of false activation. Two capacitor banks were identified from system maps. Inspection of the capacitor banks revealed an open fused disconnect on the A phase capacitor. A capacitor bank with a failed phase injects high levels of harmonic current into the system and into the earth. These harmonics caused the problem. Mitigation: The capacitor bank with the open phase was taken out of service until a troubleman could replace the blown fuse.

Figure 14-2 Fused Disconnects (Closed on Left, Open or Blown on Right)

Case Study #12 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: False activation of highway grade crossing equipment 14-16 __No_ __No_ _____ _____ _____ _____ __No_ _Yes_

Case Studies

Executive Summary: Crossing gates were activating when no trains were present. There was no pattern apparent to the false activations. The railroad had doubled their block lengths as a means of reducing maintenance costs. AC interference from parallel transmission lines had not been considered. It is now necessary to maintain track circuit balance at extremely low levels (<5%). Details: The power company personnel were familiar with the section of track due to past problems with failed insulated joints. During the initial inspection it was noted that some of the prior signal locations were out of service. It was explained that the railroad had initiated a corporate mandate to eliminate signal locations as a cost reduction measure. The effect was to double signal block lengths. This approximately doubled the rail-to-ground voltages magnetically induced by the parallel transmission lines. Because the rail-to-rail voltage that effects the grade crossing equipment is approximately equal to the rail-to-ground voltage times the track circuit unbalance (%), the railroad was forced to maintain its track circuit unbalance at half a normal level. Lessons Learned: The cost of maintaining the track circuit unbalance at unusually low levels probably outweighs any benefits of eliminating the signal locations. Case Study #13 Does continuous ac voltage at any location exceed 50 Volts? Does problem involve damage to signal equipment? Does the time of the damage correspond to the times of power system events? Did damage occur during an electrical storm? Was the dominant frequency measured 60Hz or 180Hz? Does a transmission line parallel the track? Is rail-to-rail voltage greater than 10% of rail-to-ground voltage? Is ac interference spectrum incompatible with the signal equipment present? Problem: False activation of highway grade crossing equipment Executive Summary: Starting at around 4 P.M. one day, the crossing has been intermittently down with no trains present on a more-or-less continuous basis. A shorted track circuit surge arrester caused the problem. Details: The crossing affected (Y) was using an operating frequency of 156 Hz. Measurements of the Rail-to-Ground voltages showed about 12.6 volts ac rms on both rails of the affected track. The other track showed only about 5.1 Volts ac rms rail-to-ground potential. A distribution line from a nearby substation and power plant ran immediately parallel to the tracks along the entire signal block, with 20 or less between the centerline of the track and the distribution line. The spectrum of the interference showed 60 Hz to be slightly stronger than the 180 Hz. (Zero to +3 dB stronger than the 180 Hz.) 14-17 __No_ __No_ _____ _____ _____ _____ _Yes_ _____

Case Studies

The rail-to-rail voltage on the affected track (with crossing predictor in operation) showed 2.5 volts ac rms, while the unaffected track showed only 0.6 to 0.9 volts ac rms from rail to rail. So, the rail-to-rail ac rms voltage was about 20% of the rail-to-ground voltage indicating a track circuit unbalance. The affected crossing was located in a signal block roughly 6,900 feet long, with voltages on the affected track as shown below. All voltages were measured with the crossing and other signal equipment in operation. Crossing Y was the 156 Hz crossing in trouble.
Location North I.J.s Crossing Y Crossing X South I.J.s Footage 6,917 4,594 2,481 0 Rail-Rail Volts 2.5 Vac rms 2.5 Vac rms 2.5 Vac rms 2.5 Vac rms Rail 1-Ground Volts 12.7 Vac rms 12.6 Vac rms 4.0 Vac rms 0.37 Vac rms Rail 2-Ground Volts 12.7 Vac rms 12.6 Vac rms 4.1 Vac rms 2.6 Vac rms

Inspecting the surge panel at the South insulated joints showed the following:

Figure 14-3 Shorted Surge Arrester

The arrester from the South IJ (in-line with the upper red wire) was found to measure 1.6 Ohms, providing an almost direct connection from one rail to ground. After replacing the shorted arrester, the rail-to-ground voltages at on the affected track at Crossing Y were 5.0 to 5.1 volts ac rms, much like the unaffected track at this location. Mitigation: The shorted lightning arrester was replaced.

14-18

15
PLANNING GUIDELINES FOR NEW/UPGRADED RAILROAD SYSTEMS
This chapter provides a detailed review of the design elements common to railroad system enhancements that can affect potential ac interference. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE.

15-1

Planning Guidelines for New/Upgraded Railroad Systems

Introduction
This chapter describes what ac interference issues railroad personnel should consider whenever changes are proposed to an existing joint corridor. It is assumed that the existing joint use corridor is reasonably free of ac interference problems. The point is to insure that changes to the railroad systems do not allow interference that couples into railroad systems to increase to unacceptable levels.

Quick-Start Version
Whenever changes are proposed to railroad systems it is necessary to determine if the changes are likely to increase ac interference problems or reduce ac interference problems. If the change is likely to increase ac interference, then an analysis needs to be made to determine if the increased interference is acceptable. The balance of this Handbook addresses how that analysis is done. What follows are two lists. The first list shows changes that tend to increase interference. If your proposed change is on this list, then expect to spend some time evaluating the ac interference levels that will result from the change. The second list shows changes that tend to decease interference. If your proposed change is on this list, then you will probably have reduced interference levels. Communication channels should be maintained between the railroad and the power company. Each company should notify the other of any proposed changes that could impact the level of ac interference.

15-2

Planning Guidelines for New/Upgraded Railroad Systems

Table 15-1 Changes that tend to Increase Ac Interference Levels (Bad Things) Track alignment Decreasing the horizontal distance between the track and power lines Decreasing the vertical distance between the track and power lines Any change in track alignment IF buried counterpoise was installed to mitigate magnetic induction Increasing the lengths of track blocks Operating practices Moving operating frequencies closer to power system fundamental and odd harmonics (60 Hz: 180, 300, 420, 540; 50 Hz: 150, 250, 350, 450) Adding grade crossing train detection equipment based on audio frequency track circuits Changing cab signal circuits to 60 Hz Installing wideband couplers around insulated joints Adding bandpass filters (or hardwire grounds) to balance track circuit by draining rail with higher ac interference level Connecting railroad grounds to power company grounds Removing grounds from un-signaled sidings Table 15-2 Changes that tend to Decrease Ac Interference Levels (Good Things) Track alignment Increasing the horizontal distance between the track and power lines Increasing the vertical distance between the track and power lines Reducing the lengths of track blocks Operating practices Moving operating frequencies away from power system fundamental and odd harmonics (60 Hz: 180, 300, 420, 540; 50 Hz: 150, 250, 350, 450) Changing from track circuit based grade crossing train detection equipment to systems that are not based on audio frequency track circuits Changing cab signal circuits to frequencies other than 60 Hz Inspect insulated joints regularly (use methods and criteria shown in Chapter 8) Removing wideband couplers around insulated joints Ground un-signaled sidings

15-3

Planning Guidelines for New/Upgraded Railroad Systems

Detailed Version
Whenever a change is made to a part of a system there are three possible outcomes with respect to interference. The change will have a positive effect, a negative effect, or no effect on the system. If we are changing the railroad equipment or operational practices there is the possibility of altering the characteristics of the sources, receptors, and coupling paths. It is necessary to consider all aspects of the systems involved and all types of interactions. In each case where a change might increase the level of ac interference into the receptor, an evaluation must be done to insure safety and correct operation of all systems under all reasonable operating conditions. What follows are conditions that could result from changes to railroad systems. Wherever possible the changes are described as positive (reduces ac interference coupled into the receptor) and negative (increases ac interference coupled into the receptor). Any change that has a negative impact requires some further evaluation to insure that ac interference does not exceed acceptable levels. Such further study is described elsewhere in this Handbook. Communication channels should be maintained between the railroad and the power company. Each company should notify the other of any proposed changes that could impact the level of ac interference. Sources Changes to railroad systems seldom change the sources of ac power interference except in electrified territory. However, some seemingly insignificant changes can have a surprising impact on sources of interference. For example, depending on how it is powered, a railroad battery charger can become a source of ac interference. A railroad battery charger can have a failed diode in its power supply. A failed diode generates harmonic voltages that go back through the transformer to the electric supply lines. If the charger is supplied at 120 Volts (connected between a hot wire and a grounded neutral wire), then those harmonic voltages can get into the signal systems through the ground system and cause false activations of highway grade crossing train detection equipment. But, by changing the charger power supply to be powered by 240 Volts (connected between two hot wires with no direct connection to a grounded neutral), then the harmonics generated by a failed diode go back through the electric company system without a direct connection to local ground. So, the conducted path for the ac interference can be eliminated by a relatively 1 simple change. Two other railroad signal systems that have the potential to act as sources of ac interference are 60 Hz Cab Signaling and Style C track circuits. The use of either of these circuits creates a fairly direct connection between the power company and the railroad tracks, albeit a transformerisolated one.

Note that the source may still be present (failed diode in power supply), but removing the path eliminates the ac interference. If all battery chargers were powered line-to-line, then interference from this occasional source would be effectively eliminated.

15-4

Planning Guidelines for New/Upgraded Railroad Systems

Paths Changes to railroad systems frequently change the coupling paths ac interference uses to get from sources to receptors. Distance Anything that reduces the distance between the source and the receptor tends to have a negative effect (increases ac interference in the receptor). This applies for magnetic induction, electric induction, and conduction. This is so universally understood that the only example we will give here is an exception to this rule: If a grounded counterpoise wire was buried in the earth near the track (between the power line and the track) as a mitigation measure for magnetic induction, then moving the track further from the power line could increase the amount of interference coupling into the rails, Figure 15-1. This is because you would also increase the spacing between the track and the counterpoise. When a counterpoise is placed near the receptor to mitigate magnetic induction, its effectiveness decreases with increased spacing. If the track is moved five feet, the distance from the counterpoise might double while the distance from the source power line might only increase a few percent. The result is that the effectiveness of the counterpoise would be reduced dramatically, and the interference in the receptor might increase.

Figure 15-1 Moving Track Alignment Away from the Source can Lead to Increased Interference if a Buried Counterpoise was Used to Mitigate Magnetic Induction

15-5

Planning Guidelines for New/Upgraded Railroad Systems

Block Length, Wideband Shunts, and Tuned Couplers The length of electrically continuous rail (at the interference frequency) is directly proportional to the interference voltage that will couple through magnetic induction. So, if the block length is doubled, the interference voltage will double. Pay attention to this one! Every manager wants to increase block lengths to reduce equipment and maintenance costs. Remember that there is a down side. Also, if a crossing approach has the end of a signal block inside it, the approach is sometimes extended past the insulated joints by using wideband shunts to pass the ac from the highway grade crossing train detection equipment while blocking the dc track circuit. This practice extends the electrically continuous rail length to include both blocks. It is this longer electrically continuous rail that determines the magnetically induced voltage. So, this practice of bypassing insulated joints with wideband shunts significantly increases magnetically induced interference. (It also leads to frequent filter replacement.) In ac interference environments, all coupling devices (tuned couplers and wideband shunts) used to connect one track circuit to another represent weak points in the system which may be susceptible to surges induced onto the tracks by ac power lines during a power-line fault condition. Avoid using couplers if at all possible (even though this may require the installation of additional track circuits). If couplers cannot be avoided, remember that wideband couplers are just that: wideband. Consisting essentially of a very large electrolytic capacitor (hundreds of microfarads), they are very effective at coupling track circuits together. This can increase the size of the railroad rail antenna picking up 60-Hz energy from nearby parallel ac power lines. Tuned couplers are made to be frequency selective, and will generally couple 60-Hz energy less effectively, particularly if they are tuned for a much higher frequency. However, it is often the odd harmonics of 60 Hz (180 Hz and 300 Hz in particular), which are also carried in smaller amounts on the power lines, which causes the problems. Therefore the ac signaling frequency being coupled from one block of track to the next should be chosen to avoid these power-line harmonic frequencies insofar as possible. And the use of multiple tuned couplers tuned for a range of frequencies can result in an essentially wideband coupling network being formed. It will lack the very low frequency coupling characteristics of a true wideband coupler, but can still couple the odd harmonics of 60 Hz quite effectively. Furthermore, tuned couplers may be more susceptible to damage than the more-robust wideband units. And there are much stricter application guidelines for the use of tuned couplers than there are for wideband couplers. The choice made here represents a significant trade-off. Carefully chosen tuned couplers may provide reduced ac track voltages, as compared with wideband couplers, in exchange for possibly lower reliability. Ballast The material and condition of the ballast has a big effect on the conductive coupling path. Railroad personnel know that the ballast impedance has a significant effect on the operation of track circuits. So, it wont be a big surprise that these same variations in ballast impedance can cause a wide range of interference levels at different locations and over time. Ballast impedance can vary from less than an Ohm to hundreds of Ohms. When ballast impedance is low, 15-6

Planning Guidelines for New/Upgraded Railroad Systems

conducted interference can increase. But, capacitive Induction can be reduced at low ballast impedance because the energy cant build up in the track circuit. Personnel safety in fault conditions can also be affected by variations in ballast impedance. Plan to use proper track construction techniques in areas of frequent rainfall, or other environmental conditions, which can adversely affect track ballast resistance. This means using an adequate amount of ballast rock, particularly in areas of poor drainage. This is a good practice in general, but it also has several (often unrecognized) effects on the operation of track circuits in ac interference environments. By ensuring good ballast drainage, and hence, adequate ballast resistance, the sort of one rail shorted to ground scenarios that can sometimes occur in mud-holes in the track are easily avoided. Furthermore, one of the effects of low ballast resistance on some crossing predictors is to make them behave more like motion sensors. This increases their susceptibility to ac interference. Receptors Since the receptor is usually the railroad track circuit, changes to the track circuits can change the ac interference levels. While physical changes to the track alignment and length have been covered in the path section above, there are many changes to the track circuit that can impact the levels of ac interference and how the system responds to that interference. Among the issues to be concerned about are operating frequency, circuit impedance, and circuit balance. A change in the operating frequency of railroad signal devices will affect how the devices respond to the interference present. If operating frequencies can be moved away from interference frequencies, then problems should be reduced. Figure 15-2 shows a typical frequency spectrum measured rail-to-rail.
Rail-to-Rail Voltage
0.7 0.6 0.5 Voltage (V) 0.4 0.3 0.2 0.1 0 0 60 120 180 240 300 360 420 480 Frequency (Hz)

Figure 15-2 Typical Frequency Spectrum Measured Rail-to-Rail

15-7

Planning Guidelines for New/Upgraded Railroad Systems

If the impedance of the rail circuit changes significantly, then the efficiency of magnetic induction and electric induction can be affected. High impedance communication circuits are dominated by electric induction. Low impedance track circuits are dominated by magnetic induction. So, a change from communication pole-line based circuits to track circuits can change the dominant coupling path from electric to magnetic induction. This could have a significant effect on the interference in the receptor circuit, and could certainly change the types of mitigation that would be effective. Circuit balance is critical to maintaining a receptor circuit that is highly immune to interference. Because receptor circuit unbalance converts common mode interference into differential mode interference, reducing unbalance to limit the amount of harmful differential mode interference is an important task. A degraded insulated joint or an inadvertent ground on a track lead can turn relatively low levels of common mode interference into relatively high levels of differential mode interference. The type of railroad signal equipment which is usually the most sensitive to the presence of ac interference on the track is grade crossing warning equipment. Specifically, the motion sensors and crossing predictors used to control the operation of the warning devices at a grade crossing can be fooled into thinking that there is either a train approaching, or that there has been some sort of internal or external vital failure. In either case, the only safe action for this equipment to take is to activate the warning devices, and to keep them in operation until the apparent train or failure disappears. Even though some form of train detection is needed to control the crossing, the specific choice made here has a large impact on the crossings susceptibility to ac interference. In general, Crossing Predictors are inherently much more immune to the effects of ac interference than are Motion Sensors. This is due to the different ways in which these two types of signal equipment process the voltages received from the rails. Replace older signal equipment with updated versions. The rapid advancements in electronics of the last few decades have generally improved the ac interference immunity of modern signal equipment. Many signal devices are still being asked to perform well beyond their intended service lifetimes. The continued use of outdated equipment can make a railroad signal system more susceptible to ac interference than it would be with more modern electronics. Furthermore, the signal system reliability will likely be improved just by the removal of older equipment whose failure rate has begun to rise as it approaches the end of its service lifetime. At the very least, planners of signal improvement projects should consider overhauling the existing equipment. This should be done by having it serviced by competent (usually factory-trained) technicians. Signal equipment manufacturers are often willing to arrange special terms for customers wanting to tune up a large quantity of their existing equipment. Steady-State While some changes to the power system effect only steady-state interference levels, most changes to railroad systems have the same affect whether the condition is steady state or fault. However most steady-state problems arise from differential mode interference and involve abnormal equipment operation. So, a change that increases track circuit unbalance can cause steady-state problems without affecting fault issues.

15-8

Planning Guidelines for New/Upgraded Railroad Systems

Fault While some changes to the power system effect only fault interference levels, most changes to railroad systems have the same affect whether the condition is steady state or fault. However most fault problems arise from common mode interference and result in damage to equipment or personnel electric shock hazards. It is possible to make a change that can increase fault levels without affecting steady-state issues. An example is the questionable practice of connecting a railroad ground directly to a power company ground. General Changes made to railroad signal circuits are not without significant cost. The time and personnel required to create, document, install, and test even a small change can easily run into the thousands of dollars. Larger projects can easily exceed $1M in total cost. Therefore these signal equipment changes and improvements are made only when warranted by changes in train traffic, track layout, or other external factors such as highway construction. The overall objectives outlined above should be followed to the furthest extent possible. In order to make the best use of a limited budget, be sure to contact the sales and technical representatives from your signal equipment suppliers. They are genuinely interested in satisfying their customers, and would much rather help you design a signal project in such a way as to avoid trouble, than have to help you fix it later. In many cases of ac interference, the problem arises from a seemingly insignificant design detail that could easily have been avoided, if it had been caught early enough. All too often, the signal equipment gets blamed for poor performance due to ac interference, when the layout or construction of the track is actually to blame. This statement may seem overly broad, but it is backed up by a simple and very important fact. Almost all railroad signal equipment in use today is connected to the rails in a differential fashion. That is, the voltages and currents applied to the tracks in the train detection process are applied between the rails, with the rails being of either opposite polarity (dc currents), or of opposite phase (ac currents). Furthermore, these voltages are applied in a floated fashion, with no reference to the local ground or remote earth. This is accomplished largely by transformer isolation of the signaling equipment and power supplies. The only voltages that can be sensed or measured by the signal system are from one railroad rail to the other rail on the same track. AC interference, however, tends to affect the voltage of both rails at the same time with respect to ground or remote earth potential. This causes both rails to be elevated above ground potential with very little voltage difference between them. Because the signal equipment only sees differential voltages between the rails, and cannot see this above-ground ac potential, ac induction onto the track, by itself, should not interfere with the normal operation of the railroad signal equipment. What is usually required, in order for trouble to occur, is the presence of some form of electrical unbalance between the two railroad rails. For example, although both rails of a railroad track lie almost directly on the surface of the earth, they are not considered (in the railroad world) to be 15-9

Planning Guidelines for New/Upgraded Railroad Systems

either grounded or isolated from ground. However, if we were to connect one rail directly to a good low-resistance ground in the presence of an electromagnetic field (say, from a nearby parallel power line), we would expect to observe an ac voltage present between the two rails. This happens because the grounded rail would be held at or near ground potential, and the other rail would still be floating at its induced potential. Thus, a difference in potential could be measured between the rails. If this ac voltage were to be of the proper magnitude, frequency, phase, etc. it could potentially interfere with the normal operation of the railroad signal equipment. This sort of unbalance can also be caused by railroad signal equipment that has been improperly designed, constructed, or installed. It can also be caused by equipment failures such as shorted insulated joints, or shorted lightning arresters. Avoiding mistakes during the planning stage of a signal project, which could cause such an electrical unbalance in the track, is essential to proper signal system operation. Perhaps the most general rule to remember when planning a signal system is to restrict the (electrical) length of the track circuits, so as to minimize the magnitude of the voltage that can be induced. The magnitude of the (common-mode) voltage that can be induced onto a railroad rail is proportional to the electrically continuous length of the rail exposed to the electromagnetic field (usually the field created by the power line or lines). If this means restricting the physical length of each track circuit, then do it! If this means avoiding the use of electrical coupling devices such as wideband or tuned couplers around insulated rail joints, then do it! These changes are often easy to make during the planning stages, and almost impossible to get re-done after the project has been completed. Notification Advance notice and cooperation are essential for effective practices. Both the railroad and the power company should designate an office to receive and send advance notices, and should adopt procedures to handle the notices promptly. Also, each company should designate an office to receive and send notices of trouble that might be related to the other companys facility. It is good idea to have a written arrangement that clearly states the particulars of such notifications. The arrangement should spell out the contact offices, the geographic area covered, the types of facility changes that require notice, the information to be included in the notice, the form of the notice, the actions to be taken upon notice, and any special arrangements. The arrangement should cover both advance and trouble notices. Records are necessary to isolate problems and causes. Each company should keep records of its own operating conditions as well as a record of disturbances. Each company should adopt operating instructions to be followed when abnormal conditions exist that appear to be due to interference or could cause interference. These records should be provided in the trouble notices referred to above. Of particular interest should be placed on the location and time of any disturbance event. The accurate correlation of this data between the railroad and the power company is a necessary first step to any analysis. The form in Figures 15-3 through 15-6, Electromagnetic Compatibility Information, is a useful tool for defining the information that must be collected from both parties to complete a comprehensive induction analysis. 15-10

Planning Guidelines for New/Upgraded Railroad Systems

Electromagnetic Compatibility Information


Transmission Lines Paralleling Railroad Communication and Signal Circuits I General Date Supplied ________

Power Company __________________________________ Project _________________________ Railroad Company _____________________________ Consultant _______________________ Location Description _____________________________________________________________ Designation of Lines _________ _________ _________ _________

Mile Posts (from/to) - End points of exposure are considered to be where the transmission line first comes within 200 feet of the nearest track. - (circle one: Proposed/Existing) ____/____P E ____/____P E ____/____P E ____/____P E

II

Power - Electrical Conductor Size Type ___________ ___________ ___________ ___________ ___________

Date Supplied ________ ___________ ___________ ___________ ___________ ___________ ___________

Diameter (specify units) ___________

Number of subconductors per phase and separation (specify units) ___________ ___________ ___________ ___________

Final Sag (specify units) at _____ degrees ___ (C or F) (typically 120 deg. F) ___________ ___________ ___________ ___________

Shield Wire Size Type

___________ ___________

___________ ___________ ___________

___________ ___________ ___________

___________ ___________ ___________

Diameter (specify units) ___________ Geometric Mean Radius (specify units) ___________

___________

___________

___________

Final Sag (specify units) at _____ degrees ___ (C or F) (typically 90 deg. F) ___________ ___________ ___________ ___________

AC impedance per mile at _____ degrees C (typically 25 deg. C) ___________ ___________ ___________ ___________

Figure 15-3 Electromagnetic Compatibility Information Form, Page 1 of 4

15-11

Planning Guidelines for New/Upgraded Railroad Systems

II.

Power-Electrical (Continued) Shield Wire (Continued) Grounding Continuous or Segmented (identify segments on plan drawing) ___________ Ruling Span (specify units) ___________ Resistance of structure grounds ___________ Substations Name Location Ground Resistance ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Nominal Voltage(phase-to-phase) Load (MVA) Initial Average Max. Steady State Ultimate Average Max. Steady State

___________

___________

___________

___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________

Phase Current (Amperes) Initial Average Max. Steady State Ultimate Average Max. Steady State ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Figure 15-4 Electromagnetic Compatibility Information Form, Page 2 of 4

15-12

Planning Guidelines for New/Upgraded Railroad Systems

II.

Utility-Electrical (Continued) Maximum Load Current in Amperes during single-circuit operation of multiple circuit lines: ___________ Direction of Power Flow ___________ ___________ ___________ ___________ ___________ ___________ ___________

Fault Current - Maximum for fault to ground at ends of exposure (RMS Amps. Sym.) End ____________ 1 Phase 3 Phase End ____________ 1 Phase 3 Phase ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Clearing time with Primary Relay (sec.) ___________ Maximum # of Reclosing Attempts ___________ ___________ ___________ ___________ ___________ ___________ ___________

Clearing time with Secondary Relay (sec.) ___________ ___________ ___________ ___________

Division of fault current between overhead ground wires and earth: ___________ Estimated earth resistivity Estimated length of exposure Proposed Existing ___________ ___________ ___________ ___________ Ohm-meters ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Figure 15-5 Electromagnetic Compatibility Information Form, Page 3 of 4

15-13

Planning Guidelines for New/Upgraded Railroad Systems

III

Utility - Mechanical

Date Supplied ________

Section drawings showing: Dimensions and positions of all phase wires and overhead ground wires relative to the structure centerline. (For the purpose of inductive coordination studies, conductor heights are assumed to be the attachment height minus 2/3 * final sag at temperature.) The horizontal distance between the centerline of power structures, the centerline of railroad pole line, and centerline of main track throughout the exposure. Line number and phase designation for each conductor. Elevation differences between railroad and transmission facilities. Schematic drawing of the power system in the area near the exposure showing the paralleling lines, locations of all transposition points, substations and location of all neutral grounding points throughout and near the ends of the exposure. IV Railroad Date Supplied ________

Drawings showing the railroad tracks, locations of insulating joints, the resistance between the rails, whether the rails are welded or bolted, the rail weight, and the locations of any existing ground mats or ground rods installed by railroad. Drawings showing the details of the signal circuits, their associated equipment (relays, batteries, etc.), the locations of the signal and communication circuits, whether the signal circuits are installed in open wire lines or buried, the wire size, length, and the number of all signal and communication wires, and all details regarding their physical location, the structures supporting the wires, the locations of any transpositions, and the details of any existing shielding scheme. Specific data on the existing surge protection facilities including the use of carbon blocks, gas tubes, etc. and their grounding method and resistance. Drawings showing the details of communication circuits such as telephone lines and their signal levels and frequencies. Recommended induced voltage tolerable limits if different from the AAR/EEI joint agreement. Information on the susceptibility of your signal equipment to 60 Hz voltages if such information is available. (If not, we will use values published in the technical literature for similar signal equipment.) Anticipated future plans that would effect any of the above and details as to what would change.

Figure 15-6 Electromagnetic Compatibility Information Form, Page 4 of 4

15-14

16
PLANNING GUIDELINES FOR NEW/UPGRADED POWER SYSTEMS
This chapter provides a detailed review of the design elements common to power system enhancements that can affect potential ac interference. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE.

16-1

Planning Guidelines for New/Upgraded Power Systems

Introduction
This chapter describes what ac interference issues power company personnel should consider whenever changes are proposed to an existing joint corridor. It is assumed that the existing joint use corridor is reasonably free of ac interference problems. The point is to insure that changes to the power systems do not allow interference that couples into railroad systems to increase to unacceptable levels.

Quick-Start Version
Whenever changes are proposed to power systems it is necessary to determine if the changes are likely to increase ac interference problems or reduce ac interference problems. If the change is likely to increase ac interference, then an analysis needs to be made to determine if the increased interference is acceptable. The balance of this Handbook addresses how that analysis is done. What follows are two lists. The first list is changes that tend to increase interference. If your proposed change is on this list, then expect to spend some time evaluating the ac interference levels that will result from the change. The second list is changes that tend to decease interference. If your proposed change is on this list, then you will probably have reduced interference levels. Communication channels should be maintained between the power company and the railroad. Each company should notify the other of any proposed changes that could impact the level of ac interference.

16-2

Planning Guidelines for New/Upgraded Power Systems

Table 16-1 Changes that tend to Increase AC Interference Levels (Bad Things) Power line alignment Decreasing the horizontal distance between the track and power lines Decreasing the vertical distance between the track and power lines Increasing the length of the parallel exposure Increasing the spacing between phase wires Equipment Increasing the current carrying capacity of the lines (reconductoring, retensioning, restringing, rerating, etc.) Increasing the fault current potential (added generation, etc.) Removing fault current restricting devices (inductors, etc.) Using undersized neutral wires Operating practices Multiple reclosings on faults Manual reclosing on faults Table 16-2 Changes that tend to Decrease AC Interference Levels (Good Things) Power Line Alignment Increasing the horizontal distance between the track and power lines Increasing the vertical distance between the track and power lines Phasing multiple circuits for minimum magnetic field levels at rails Decreasing the length of the parallel exposure Decreasing the spacing between phase wires Adding a second circuit to a single circuit corridor Equipment Decreasing the current carrying capacity of the lines (de-rating) Install fault current limiting devices Using oversized neutral wires Avoid equipment that can produce low order harmonics Operating practices Reduce the number of reclosings on faults for line parallel to tracks Inspect capacitor banks regularly

16-3

Planning Guidelines for New/Upgraded Power Systems

Detailed Version
Whenever a change is made to a part of a system there are three possible outcomes with respect to interference. The change will have a positive effect, a negative effect, or no effect on the system. If we are changing power system equipment or operational practices there is the possibility of altering the characteristics of the sources and coupling paths. It is necessary to consider all aspects of the systems involved and all types of interactions. In each case where a change might increase the level of ac interference into the receptor, an evaluation must be done to insure safety and correct operation of all systems under all reasonable operating conditions. What follows are conditions that could result from changes to power systems. Wherever possible the changes are described as positive (reduces ac interference coupled into the receptor) and negative (increases ac interference coupled into the receptor). Any change that has a negative impact requires some further evaluation to insure that ac interference does not exceed acceptable levels. Such further study is described elsewhere in this Handbook. Communication channels should be maintained between the power company and the railroad. Each company should notify the other of any proposed changes that could impact the level of ac interference. Sources Anything that carries current is a source of magnetic fields, and a possible source of magnetic induction. Anything that has a voltage on it is a source of electric fields, and a possible source of electric induction. Any electrically energized object with a connection to earth or a conductive path to a receptor is a possible source of conducted interference. So, within the power industry almost everything is a possible source of ac interference. It should also be noted that many things outside the power company could be sources. Most of these outside sources of ac interference are customers of the power company that have less than ideal wiring and facilities. Unfortunately, in must places, the power company has little or no authority over these third-party sources. What follows is a listing of variables within the power system. Each is followed by a short discussion regarding that variables effect on the power system as a source of ac interference. Voltage The voltage of the power lines has a direct effect on the electric field level, and therefore, electric induction. Higher voltage has a tendency to increase electric field levels. When open wire communication lines were more common on railroads, this was often the driving concern for inductive coordination. Because communication circuits are high impedance receptors, electric induction is the dominant coupling path. But, with the elimination of most open wire communications, the track circuits became the driving concern. These low impedance receptors are more susceptible to magnetic induction, which is current driven.

16-4

Planning Guidelines for New/Upgraded Power Systems

So, while any study of a joint corridor should still include the effects of electric induction, the voltage of the power line is usually not the direct concern. While there is sometimes a correlation between higher voltage and higher current, sometimes there is no correlation. It is not possible to make a blanket statement that any change in voltage will have a specific effect (good or bad). Current Steady State Increased steady state current levels in a power line result in increased magnetic induction. While load currents vary continuously, the highest likely levels are used to evaluate magnetic induction. It would be unacceptable for railroad signaling to stop working every time power company loads peaked. If a corridor was designed for peaceful coexistence under expected loads, and the power company then increased the possible peak loading on the line, there is go assurance that the magnetically induced energy could be tolerated. Increases in steady state current require further study. Current Fault and Switching Short duration elevated current on power lines usually result from faults and switching operations. Switching operations are shorter duration event and usually a lesser concern, so faults are analyzed. The worst case for fault current is typically a bolted fault (0 ohms) on the phase conductor closest to the rails with the location of the fault near the end of the railroad parallel. Every power line, at every point, has a fault current potential. That is the amount of current that would flow in the worst-case fault. Both the magnetic induction and the local earth conduction from such faults present the potential for shock hazard and equipment damage. Sometimes changes in the power grid change the fault current potential. The addition of a generating station (such as a gas turbine unit) could increase the fault current available. Increases in available fault current require further study. Phasing The relative phasing of multiple circuit rights-of-way can have a significant effect on steady state magnetic induction. (It should be noted that phasing has absolutely no effect on fault induction.) By carefully arranging the phase conductors of multiple circuits (or split phasing a single circuit) the magnetic field levels and magnetic induction can be significantly reduced. Unfortunately, it is prudent to consider any joint corridor with any one circuit off line for maintenance. This sometimes substantially reduces the benefit of phase cancellation. Changes in phasing sometimes require further study. Balance Much has been made of the benefits of balancing the current on the three phases of a power line. And, to the extent practical, lines should be balance by the power company. Unfortunately, the load on any power line varies continuously, and is controlled by the customers not the power company. The best any power company can do is to move load from phase to phase to achieve 16-5

Planning Guidelines for New/Upgraded Power Systems

the best possible balance on a typical day. The problem comes when customers turn things on and off. So, for any power line some amount of unbalance is inevitable. The amount of variation in phase currents and balance can be relatively large particularly for distribution lines. In short, you cannot count on phase balancing to solve most ac interference problems. Transmission A transmission line parallel of a railroad track is the classic example of inductive coordination with which most power company and railroad engineers are most familiar. Chapters 11 and 12 of this handbook provide in depth study into the issues. Knowledge of the impacts of various aspects of transmission line design that effect EMI on the railroad must be incorporated early in the siting/design process. Chapter 11 can be particularly helpful. Distribution Distribution lines usually carry substantially less current than transmission lines. But, there are two aspects of distribution lines that can be of concern. First, distribution lines often have strong harmonic levels. On a 60 Hz system, 180 Hz and 300 Hz levels can be relatively high. Because these harmonics are near the operating frequencies used by various audio frequency railroad signal equipment, the potential for operational problems exists. Second, if the distribution system uses a multigrounded neutral, then the harmonic rich neutral current also flows in the earth. This introduces earth conduction as a possible path for EMI. This is often the combination that contributes to false activation of grade crossing train detection equipment. Increasing the size of the neutral conductor (reducing the impedance) will help reduce the proportion of neutral current in the earth. Paths Changes to power systems frequently change the coupling paths that ac interference uses to get from the sources to the receptors. Magnetic Induction Anything that reduces the distance between the source and the receptor tends to have a negative effect (increases ac interference in the receptor). This applies for magnetic induction, electric induction, and conduction. Magnetic induction is the dominant coupling path into track circuits. Magnetic induction is usually associated with transmission lines. But, distribution lines can also couple through magnetic induction.

16-6

Planning Guidelines for New/Upgraded Power Systems

Electric Induction Electric induction is the dominant coupling path into high impedance communication circuits. Electric induction is often an issue when transmission lines parallel a railroad with open wire communication circuits. Earth Conduction Earth conduction is usually only a steady state issue where distribution lines have multigrounded neutrals. The harmonic rich neutral currents also flow in the earth. From the earth they couple through the ballast into the track circuits. Levels of ac interference through this path tend to be highest when the source substation is near the tracks. Earth conduction is also an issue whenever faults occur. The ground potential rise is most pronounced at the site of the fault with a voltage gradient toward remote earth in all directions. Such conducted energy is a potential step and touch hazard near the fault. If railroad equipment is within the area of ground potential rise, equipment damage is possible, and elevated voltages could be carried on railroad conductors. Receptors The receptor is usually the railroad track circuit. A reasonable measure to use when evaluating the source/path combination from power lines is the rail-to-ground voltage per unit length. Rail-to-ground voltage is roughly proportional to the track circuit length. And, rail-to-rail voltage is primarily dependent on track circuit balance. So, measures beyond rail-to-ground voltage per unit length are beyond the power companys direct control. Steady-State Any change to the power system that can increase steady-state rail-to-ground voltages should be studied. Such changes could increase the likelihood of equipment operation issues and personnel safety. Fault Most fault related problems arise from common mode interference and result in damage to equipment or personnel electric shock hazards. It is possible to make a change that can increase fault levels without affecting steady-state issues. A change to the transmission grid at a remote location that could increase fault current available would require further study. Notification Advance notice and cooperation are essential for effective practices. Both the railroad and the power company should designate an office to receive and send advance notices, and should adopt procedures to handle the notices promptly. Also, each company should designate an office to receive and send notices of trouble that might be related to the other companys facility. 16-7

Planning Guidelines for New/Upgraded Power Systems

It is good idea to have a written arrangement that clearly states the particulars of such notifications. The arrangement should spell out the contact offices, the geographic area covered, the types of facility changes that require notice, the information to be included in the notice, the form of the notice, the actions to be taken upon notice, and any special arrangements. The arrangement should cover both advance and trouble notices. Records are necessary to isolate problems and causes. Each company should keep records of its own operating conditions as well as a record of disturbances. Each company should adopt operating instructions to be followed when abnormal conditions exist that appear to be due to interference or could cause interference. These records should be provided in the trouble notices referred to above. Of particular interest should be placed on the location and time of any disturbance event. The accurate correlation of this data between the railroad and the power company is a necessary first step to any analysis. The form in Figures 16-1 through 16-4, Inductive Coordination Information, is a useful tool for defining the information that must be collected from both parties to complete a comprehensive induction analysis.

16-8

Planning Guidelines for New/Upgraded Power Systems

Electromagnetic Compatibility Information


Transmission Lines Paralleling Railroad Communication and Signal Circuits I General Date Supplied ________

Power Company __________________________________ Project _________________________ Railroad Company _____________________________ Consultant _______________________ Location Description _____________________________________________________________ Designation of Lines _________ _________ _________ _________

Mile Posts (from/to) - End points of exposure are considered to be where the transmission line first comes within 200 feet of the nearest track. - (circle one: Proposed/Existing) ____/____P E ____/____P E ____/____P E ____/____P E

II

Power - Electrical Conductor Size Type ___________ ___________ ___________ ___________ ___________

Date Supplied ________ ___________ ___________ ___________ ___________ ___________ ___________

Diameter (specify units) ___________

Number of subconductors per phase and separation (specify units) ___________ ___________ ___________ ___________

Final Sag (specify units) at _____ degrees ___ (C or F) (typically 120 deg. F) ___________ ___________ ___________ ___________

Shield Wire Size Type

___________ ___________

___________ ___________ ___________

___________ ___________ ___________

___________ ___________ ___________

Diameter (specify units) ___________ Geometric Mean Radius (specify units) ___________

___________

___________

___________

Final Sag (specify units) at _____ degrees ___ (C or F) (typically 90 deg. F) ___________ ___________ ___________ ___________

AC impedance per mile at _____ degrees C (typically 25 deg. C) ___________ ___________ ___________ ___________

Figure 16-1 Electromagnetic Compatibility Information Form, Page 1 of 4

16-9

Planning Guidelines for New/Upgraded Power Systems

II.

Power-Electrical (Continued) Shield Wire (Continued) Grounding Continuous or Segmented (identify segments on plan drawing) ___________ Ruling Span (specify units) ___________ Resistance of structure grounds ___________ Substations Name Location Ground Resistance ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Nominal Voltage(phase-to-phase) Load (MVA) Initial Average Max. Steady State Ultimate Average Max. Steady State

___________

___________

___________

___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________

___________ ___________ ___________ ___________

Phase Current (Amperes) Initial Average Max. Steady State Ultimate Average Max. Steady State ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Figure 16-2 Electromagnetic Compatibility Information Form, Page 2 of 4

16-10

Planning Guidelines for New/Upgraded Power Systems

II.

Utility-Electrical (Continued) Maximum Load Current in Amperes during single-circuit operation of multiple circuit lines: ___________ Direction of Power Flow ___________ ___________ ___________ ___________ ___________ ___________ ___________

Fault Current - Maximum for fault to ground at ends of exposure (RMS Amps. Sym.) End ____________ 1 Phase 3 Phase End ____________ 1 Phase 3 Phase ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Clearing time with Primary Relay (sec.) ___________ Maximum # of Reclosing Attempts ___________ ___________ ___________ ___________ ___________ ___________ ___________

Clearing time with Secondary Relay (sec.) ___________ ___________ ___________ ___________

Division of fault current between overhead ground wires and earth: ___________ Estimated earth resistivity Estimated length of exposure Proposed Existing ___________ ___________ ___________ ___________ Ohm-meters ___________ ___________ ___________ ___________ ___________ ___________ ___________ ___________

Figure 16-3 Electromagnetic Compatibility Information Form, Page 3 of 4

16-11

Planning Guidelines for New/Upgraded Power Systems


III Utility - Mechanical Date Supplied ________

Section drawings showing: Dimensions and positions of all phase wires and overhead ground wires relative to the structure centerline. (For the purpose of inductive coordination studies, conductor heights are assumed to be the attachment height minus 2/3 * final sag at temperature.) The horizontal distance between the centerline of power structures, the centerline of railroad pole line, and centerline of main track throughout the exposure. Line number and phase designation for each conductor. Elevation differences between railroad and transmission facilities. Schematic drawing of the power system in the area near the exposure showing the paralleling lines, locations of all transposition points, substations and location of all neutral grounding points throughout and near the ends of the exposure. IV Railroad Date Supplied ________

Drawings showing the railroad tracks, locations of insulating joints, the resistance between the rails, whether the rails are welded or bolted, the rail weight, and the locations of any existing ground mats or ground rods installed by railroad. Drawings showing the details of the signal circuits, their associated equipment (relays, batteries, etc.), the locations of the signal and communication circuits, whether the signal circuits are installed in open wire lines or buried, the wire size, length, and the number of all signal and communication wires, and all details regarding their physical location, the structures supporting the wires, the locations of any transpositions, and the details of any existing shielding scheme. Specific data on the existing surge protection facilities including the use of carbon blocks, gas tubes, etc. and their grounding method and resistance. Drawings showing the details of communication circuits such as telephone lines and their signal levels and frequencies. Recommended induced voltage tolerable limits if different from the AAR/EEI joint agreement. Information on the susceptibility of your signal equipment to 60 Hz voltages if such information is available. (If not, we will use values published in the technical literature for similar signal equipment.) Anticipated future plans that would effect any of the above and details as to what would change.

Figure 16-4 Electromagnetic Compatibility Information Form, Page 4 of 4

16-12

17
CONCLUSIONS
This chapter provides a review of the state of affairs regarding electromagnetic effects between power lines and railroads. This includes unresolved issues and recommended future research. Brian Cramer is the Technical Manager for Electromagnetic Compatibility in the Power Delivery group of EPRI. He earned his Bachelors degree in Electrical Engineering from Lehigh University in Bethlehem, Pennsylvania. He worked for AIL Systems as an integration lab manager, with responsibility for all B-1 bomber electronic countermeasures systems. He later worked for Lockheed Electronics performing research and development on advanced phased-array antenna systems. In 1990, he joined Commonwealth Edison (Exelon) in Chicago, where his responsibilities included right-of-way and site selection, increased transmission capacity, EMF management, and electromagnetic compatibility (EMC). Later, as Technical Expert for Inductive Coordination and Electrical Effects at ComEd, he was responsible for maintaining effective working relationships with the two dozen railroads within ComEds service territory. In 2001, he joined EPRI to manage the recently formed EMC Program. In this capacity he is responsible for research into issues ranging from interference with Global Positioning Systems (GPS), railroads, and pipelines, to broadband data transmission over power lines (BPL). Other projects include the effects of High Altitude Electromagnetic Pulse (HEMP) and Intentional Electromagnetic Interference (IEMI) on power systems. He is a member of IEEE, ANSI, AREMA, and CIGRE.

17-1

Conclusions

This electromagnetic compatibility handbook is the first step in a greater process. The purpose of this process is to change the way railroad and power company people think about ac interference. Using the tools of electromagnetic compatibility, ac interference problems can be relegated to the past. Through proper application of reasonable achievable standards, tracking down an ac interference problem could be as simple making an observation or measurement, and then fixing the broken component that caused the abnormal condition. In the future being described, ac interference simply wont happen to the degree it does today. What comes next is up to the industries involved to imagine, design, and create. A good first step is to find out who the people are at the other industry that you can partner with to address these issues. Establish communication with them. Meet with then along the right-of-way. Determine what information about the railroad systems is important. Make sure you know where that information is and how to get it. Think about the records from the electric company that you will need to address steady state, switching, and fault issues. Again, make sure you know where that information is and how to get it. Think. Brainstorm about what has happened in the past, and what might happen in the future. Consider how you can best support each other. Make sure you know how to make contact in an emergency. And, read this handbook.

Current State of Affairs


We have the ability, today, to address most of the issues that arise from ac interference. Examples of issues that can usually be mitigated at reasonable cost are: Personnel safety (step and touch) under steady-state conditions Operational impacts to train control track circuits from steady-state conditions Damage to railroad signal equipment from power system switching and fault magnetic induction - with the exception of surge protective devices (SPDs, a.k.a. lightning arresters)

Issues that sometimes defy satisfactory mitigation are: Personnel safety (step and touch) under power system fault conditions satisfactory mitigation is only a problem if SPDs have been destroyed by a prior event. Highway grade crossing train detection equipment false activation Cab signaling equipment operated at 60 Hz Damage to SPDs from power system fault magnetic induction

The use of this handbook can assist in the mitigation of ALL of these potential issues. But, as indicated, some issues may defy satisfactory mitigation without unreasonable impacts to system operation or cost.

Short-Term Solutions
Our short-term solutions consist primarily of educating ourselves about EMC, working as partners toward effective answers to specific problems, and sharing information throughout the 17-2

Conclusions

industries involved. This handbook is one tool in this process. The handbook is intended to be a living document. So, any comments, corrections, additions, or enhancements will be gratefully accepted and incorporated into future editions.

Long-Term Solutions
Long-term solutions would result in available equipment immunity always exceeding the interference environment. Such conditions might come from grass roots efforts, but usually only follow from the imposition of standards. The ideal solution is to place an achievable limit on the energy a power company can impress on the track (longitudinal electromotive force). We would also have an achievable maximum unbalance specified for the track circuit. Together these would limit the rail-to-rail ac interference voltage on the track to a fixed level. The railroad signaling equipment would then have an immunity level that is specified at a higher interference level than is present on the track. The result would be that ac interference would never cause operational problems. Highway Grade Crossing Train Detection Equipment False Activation and Cab Signaling Equipment Operated at 60 Hz In many locations false activations related to ac power interference are a significant problem. Interference levels do not exceed existing standards. Equipment meets the existing immunity standards at some frequencies and power levels. Yet, in some situations, reliable performance is difficult to obtain and cannot be maintained over time1. For guidelines or standards to be effective, they must be structured so that the measurable quantity is controllable by the organization being held responsible. Because rail-to-rail voltage is also a function of block length and track circuit balance, it is difficult to make rail-to-rail voltage the responsibility of the power company. A quantity used in some EMC standards is the longitudinal electromotive force (LEF, a.k.a. longitudinal electric field). This is a measure of the ability of the power lines magnetic field to induce a voltage in a parallel conductor. LEF has units of volts per unit length. It would then fall to the railroad to decide on block lengths based on tradeoffs of advantages of longer blocks versus increased rail-to-ground interference. The rail-to-rail interference resulting from the rail-to-ground level is almost entirely related to track circuit balance. So, it would be the railroads responsibility to maintain the needed track circuit balance. And, finally, the equipment used to detect the trains needs to be less susceptible to ac interference. Standards of immunity need to be extended to harmonics of the power frequency fundamental. The questions that need to be addressed by the industries are:
1

Cab signaling equipment operating at 60 Hz is lumped in with highway grade crossing equipment because both kinds of systems are susceptible to the same relatively low levels of differential ac interference.

17-3

Conclusions

Is the above distribution of responsibilities reasonable and appropriate? If not, what should they be? What level of track circuit balance can be reasonably required of railroads? How should track circuit balance be measured? What levels of longitudinal electromotive force can be reasonably required of the power industry? Is longitudinal electromotive force a reasonable measure if the interference is coupled into the track by earth conduction? What other measure should be used? How should longitudinal electromotive force be measured? Are there ways to use existing technologies for constant warning time devices differently to reduce false activations? Can the Committees streamline the approval process from the FRA for such non-standard applications? What are the most promising proposed technologies to replace existing constant warning time devices? What can AREMA do to accelerate the development and approval of successful technologies? What levels of ac interference immunity can be reasonably required of constant warning time device suppliers? What levels of ac interference immunity should be required of new technologies?

Damage to SPDs from Power System Fault Magnetic Induction Regarding the failure of SPDs under power system fault conditions, the first question to answer is whether or not there is a problem. Track circuit SPDs are designed to withstand a standard lightning waveform. Extreme lightning will destroy the SPDs on occasion. Magnetic induction from power system faults will destroy track circuit SPDs easily. This is a potential problem for two reasons: Once the SPD is gone (open), part of the protection against step and touch voltages is gone resulting in increased risk to personnel. Once the SPD is gone (open), part of the protection against equipment damage is gone resulting in increase risk of equipment damage.

SPDs are expected to be sacrificial. They are designed to fail in such a way that the equipment is protected. In most systems, once the arrester is gone, the track wire is still connected to the equipment inside the enclosure. Some railroads and some equipment manufacturers are installing fuses or fused links of wire in the track circuit leads just before the SPDs. The effect of these fuses is that the track circuit is disconnected from the equipment in the enclosure if excess voltage is present. The down side is that the track signals will be off-line after the fuse is open. The advantage is the subsequent potentially damaging voltages will be kept outside the

17-4

Conclusions

enclosure and away from personnel and equipment. These fuses can be designed to protect the arresters as well. In cases where fusing the track circuit is unacceptable, another long-term solution needs to be found. One such solution might involve a new kind of SPD that can handle the energy of magnetically induced power system fault voltages while meeting the other requirements of vital track circuit arresters.

Future Research
In addition to the above-mentioned needs, there are other issues that deserve further study. One example of such a need is insulated rail joints. Specifically, there have been some observations that insulated joints seem to fail more often in track that parallels heavily loaded power transmission lines. Research into this possibility would have three components: Surveys to determine if the association is real Studies of failed joints to determine the failure mechanism(s) Recreation of failures in a controlled environment

Once these issues are resolved, design work can begin of ways to mitigate the problem. A second area of study would support the growing requests to install dc transmission along railroad rights-of-way. A detailed evaluation of the effects of such a shared corridor has never been published in the available literature. While some aspects have been studied, a complete analysis that could be applied to any proposed situation would be helpful.

Conclusion
While it has been over one hundred years in the making, our work has just begun. With careful application of the processes in this handbook, we should be able to effectively mitigate most problems. With future research and development we should be able to create an environment where the equipment will always work together. And, with proper vigilance we should be able to maintain harmony as new technologies and equipment continue to improve our ability to deliver freight, passengers, and electric power.

17-5

18
GLOSSARY
This chapter provides a comprehensive glossary of railroad and power terms, because the first step toward amicable coexistence is effective communication. Mr. Silva is the President of Enertech Consultants. He holds a BS and MS in Engineering in 1971 and 1976, respectively. Mr. Silva has a state of California professional engineering license in Electrical Engineering and is also a registered professional engineer in seven other states. He has 35 years of experience related to electric power facilities, electric and magnetic fields (EMF), and applied research projects. Early in his career, at the Southern Company, he supervised an electric transmission line design group and has worked at the Electric Power Research Institute where he manages research on electromagnetic compatibility (EMC) and RF and Wireless technology. In 1982, he founded Enertech Consultants, a company focused on applied research on Electric and Magnetic Field (EMF), EMF instrumentation and software development, and scientific consulting. Enertech designs, manufacturers, and sells EMF instrumentation around the world in 46 countries and has probably performed more EMF studies than any company in the United States. Large projects include: the Electric Power Research Institute, United Nations, U.S. Department of Energy, and California Department of Health Services. He has authored numerous articles and papers on power frequency electric and magnetic field exposure assessment. He has recently published a new study of the use of the Global Positioning System near to electric power facilities and has coauthored the 500 page Electric and Magnetic Field Management Reference Book and CD-ROM. He is a Senior Member of IEEE, and a member of BEMS and the Institute of Navigation. Michael R. House earned his Bachelors degree in Electrical Engineering from California State Polytechnic University in Pomona, and worked for General Dynamics R.F. Guidance Design Group for six years, followed by more than twelve years at Safetran Systems Electronic Division in California. While at Safetran he worked in the Technical Support Department on a variety of projects, including in-house and field research into ac power interference problems, an investigation into poor shunting with the Association of American Railroads, the design of train-motion simulator systems for motion sensors and crossing predictors, the development and delivery of customer training programs including portable track simulators, and the resolution of specific customer application problems with crossing warning systems. He is married to a woman who actually understands why he takes pictures of power lines and railroad crossings while on vacation, and now does independent consulting work for Timerider Technologies, Inc., with a diverse base of customers including railroads, electric utilities, and governmental agencies. He is a member of IEEE and AREMA.

18-1

Glossary

Introduction
A comprehensive set of railroad and power terms is integrated into a single glossary to assist railroad and power engineers using this handbook. Most of the railroad definitions derive from four main sources: AREMA Signal Manual Part 1.1.1 (Recommended Definitions for Technical Terms Used in Railway Signaling, Revised 1997); Railway Technical Web Pages, Glossary of Modern Railway Terminology (TrainWeb.com); AREMA Signal Manual, Part 11.3.10, (Purpose and Meaning of Terms Used in Surge Protection and Grounding, Revised 1999); and The Practical Guide to Railway Engineering, AREMA, 2003. Electric power terms were developed using the Electric and Magnetic Field Management Reference Book, EPRI, 1999 and with guidance from the IEEE Standard Dictionary of Electrical and Electronics Terms, 1996. This glossary provides practical explanations while avoiding arcane language or complex mathematics and does not emphasize a rigorous treatment for each term. Glossary terms used in railway signaling consist of two types of terms: 1) electrical, mechanical and operational terms which are current to the railway signal industry and in general railroad use and, 2) historical railway terms which are obsolete or that no longer represent recommended practice, but are still in use or are historically significant. Terms having North American and UK (or European) variations are labeled as UK where definitions may differ. Entries that are phrases rather than single words appear in their most commonly used word order, e.g., acknowledging whistle rather than whistle, acknowledging. However, similar or related terms are grouped together when there are several varieties of the same thing (i.e., usually a noun qualified by a number of different adjectives). Generally these are listed under a common heading using the noun. For the same reason, the alphabetic arrangement considers the second word of a phrase only after all terms using a common first word have been listed. For many entries, the same term has different meanings in different contexts. The intended context is shown in parentheses after the term, e.g., Back Light (Signal) or Switch (Track). When a railroad term may be confused with an electric power term, the expression (Railroad) is used. Two forms of cross-references are used: See... indicates the entry under which the definition will be found, and See also indicates a reference to a similar or easily confused term, or to a related or contrasting term that may help in understanding or limiting the definition provided.

Glossary
AAR Association of American Railroads: An organization of railroads serving the United States, Canada, and Mexico for the purpose of improving transportation services. ABS (or AB): Automatic Block Signal System. Absolute Block System: The railway signaling system based on the principle of dividing a line into sections or blocks. Normally, no train is allowed to enter a section occupied by another. 18-2

Glossary

Absolute Permissive Block System (APB): A block signal system where the block is usually from siding to siding for opposing movements and the fixed signals governing entrance into the block display an aspect indicating Stop when the block is occupied by an opposing train. For following movements, the section between sidings is divided into two or more blocks and train movements into these blocks, except the first one, are governed by intermediate fixed signals, cab signals, or both. The intermediate fixed signals usually display an aspect indicating Stop; then Proceed at Restricted Speed, and the cab signal displays an aspect indicating Proceed at Restricted Speed', as its most restrictive indication. (See also: Block). Absolute Signal: (See Signal, Application). Absorption: The attenuation of an electromagnetic wave as it travels in a material by scattering the photons of the wave with electrons of the material (so-called free-carrier absorption), also called diffusion. A photon gives up its energy to a materials electron and energy is divided in the material in this manner. Thus energy is lost from the electromagnetic wave, the end result being that the wave is weakened. The extent to which a metallic medium attenuates an electromagnetic wave by absorption is described by its skin depth - the depth in the material at which the induced current density declines to 1/e (about 37%) of its value with respect to a given point (usually at the surface where the wave enters the medium). Absorption is a function of the material and the frequency of the electromagnetic wave. (See also Skin Depth). Absorption Loss: Absorption losses are a function of the physical characteristics of a shield. For a given thickness, magnetic materials (such as steel) provide higher absorption losses than nonmagnetic materials (such as copper). When reflection losses are low, thicker highpermeability material is employed to increase shielding efficiency. In general, absorption losses are low at low frequencies and rise gradually to high levels at high frequencies. (See also Absorption). Acceptance: In UK railway signaling terms, acceptance means the permission given by a signalman for a train to enter the section of line the signalman controls. Acknowledging: (Railroad) Operation by enginemen of acknowledging device. Acknowledging Device: A manually operated electric switch or pneumatic valve by means of which, on a locomotive equipped with automatic train stop or train control device, an automatic brake application can be forestalled, or by means of which, on a locomotive equipped with an automatic cab signal device, the sounding of the cab indicator can be silenced. Acknowledging Time: (Intermittent Automatic Train Stop System) A predetermined time within which an automatic brake application may be forestalled by means of the acknowledging device. Acknowledging Whistle: An air-operated whistle that is sounded when the acknowledging device is operated or when a more restrictive cab signal is displayed. AAAC: All Aluminum Alloy Conductor. AAC: All Aluminum Conductor. 18-3

Glossary

AACSR: Aluminum Alloy Conductor Steel Reinforced. ACAR: Aluminum Conductor Alloy Reinforced. ACSR: Aluminum Conductor Steel Reinforced. A power line conductor with strands of aluminum wire wound around a core of high-strength steel wires. Often the steel core wires are galvanized (coated with zinc) for corrosion resistance. ACSS: Aluminum Conductor, Steel Supported. ACSS/TW Aluminum Conductor, Steel Supported, Trapezoidal Shaped Aluminum Strands. AFO Audio Frequency Overlay: The Audio-Frequency Overlay track circuit operates in the same manner as its dc and power-frequency counterparts. The term overlay implies that the AFO circuit is overlaid upon some other form of track circuit, although it can be used in a standalone fashion. This is also true of other types of circuits not classically defined as AFO (e.g. Crossing Predictors and Motion Sensors). In practice, a half dozen or more overlaid track circuits of various types may be present at some complex highway crossings. One of the most common applications of AFOs is for highway crossing train detection on electrified and diesel-electric railroads. Although they do not provide the operational benefits of Crossing Predictors and Motion Sensors, such as minimizing highway traffic delays, AFO track circuits of up to 4,000 feet are common in diesel-electric territory. However, the track circuit length (approach circuit) needed to provide adequate warning time for grade crossings averages 2,500 feet or less where train speeds do not exceed 50 miles-per-hour. The vital control logic used at such grade crossings usually incorporates timers and other additional watchdogs to prevent the misoperation of the grade crossing warning system due to failed AFO track circuits. AFO transmitters and receivers are generally located in close proximity (less than 100 feet) to their connections to the rails, but this can be extended for some circuits with the use of special coupling circuits. These coupling circuits enable the electronics to be located many hundreds or even thousands of feet away from where their track wires are attached to the rails. Active Loops: Active loops is a term describing a system of conductors arranged in a path (loop) and supplied an electric current by an external source for the purpose of magnetic field reduction within a region. The geometry of the loop will depend on the magnetic field source geometry. To be effective, an active loop system would utilize an active feedback control system to adjust loop current as the magnetic field within the region of interest changes. (See also Cancellation Loops). Active Shielding: Magnetic fields can be reduced (shielded) by establishing currents in wires such that the fields produced by those currents oppose the fields to be reduced. Active shielding (or shielding with active conductors) refers to any scheme to reduce the magnetic field in certain regions of space by use of conductors with an imposed current whose magnitude, direction, and phase angle create fields in opposition to the ambient fields and thereby reduce the overall magnetic field in a region. AIEE: (Historical) American Institute of Electrical Engineers. (Now IEEE). Air Brake: This is the most common type of train brake using compressed air.

18-4

Glossary

Air Gap- Magnetic: The space occupied by air between two adjacent parts of a magnetic circuit prior to the application of a protective coating. The actual physical spacing between two adjacent parts a magnetic circuit should be reduced by the thickness of any layers of non-magnetic protective coating materials. Alarm: (Overheated Journal) (Railroad). Absolute: An alarm that occurs when either end of an axle exceeds a predetermined heat level. Differential: An alarm that occurs when the heat difference between opposite ends of the same axle exceeds a predetermined heat level. Ratio: An alarm that occurs at a predetermined heat level ratio between ends of the same axle, e.g., 3 to 1 ratio. Alumoweld: Aluminum clad high-strength steel wire. American Standard: (See ANSI). Ampacity: The current, in amperes, that a conductor can carry continuously under specified conditions of use without exceeding its temperature rating. Ampere-Turn: A term in the MKS (SI) system used to describe the magnetomotive force (mmf) in a closed ring solenoid of uniform cross section. The ampere-turn is the product of the number of turns of wire (enclosed in a closed integration path) and the electric current in each turn; it is the -1 magnetomotive force (mmf). (The CGS unit of mmf is the gilbert: 4 x 10 gilbert = 1 ampereturn.) The ampere-turn per meter (mmf per unit length) is the magnetic field strength unit in MKS (SI) system. Ampres Law: Ampres Law, including an important extension of it made later by Maxwell, is one of the basic equations of electromagnetic field theory. This law equates the magnetic field integrated around a closed contour or loop to the current through the loop (or integral of the current density over a surface bounded by the loop). Ampres Law describes the connection between the field and the current:

B dl = I enclosed
At first glance one would think that the law is used to determine the current, I, by integration (indeed, this principle is how a clamp-on ammeter works to measure the current in a conductor). In practice, the current is usually known and the law provides a method of finding the magnetic field (for more complex problems, the Biot and Savart Law is used). This equation above is true, in general, for any magnetic field configuration, for any assembly of currents, and for any path of integration. (For frequencies much much higher than 60 Hz, the equation must be modified if a time-varying electric field is present within the path of integration). (See also Maxwells Equations and Quasistatic).

18-5

Glossary

Angular Frequency: The angle (radian) of a sinusoidal quantity per unit of time (second). A quantity that varies sinusoidally with time, q = Q Sin (t+) has an angular frequency equal to and has a frequency of: f = /2. ANSI: American National Standards Institute (Formerly American Standards Association). APB: Absolute permissive block. Application Unit: (See also Stepper). Control Office: A group of relays and circuits in a code system which in conjunction with the control levers forms the control code characters, and which stores the information received from indication codes. Field: A group of relays and circuits in a code system that applies control code information to the proper function control relays and assists in the transmission of indication codes. Approach Circuit: A circuit generally used in connection with announcing the approach of trains at a block or interlocking station or to provide initial activation of a highway crossing warning system. Approach Clearing Circuit: A term applied to a circuit used in connection with the operation of a signal in advance of an approaching train. Approach Control: (See Route Signaling). Approach Indicator: An indicator used to indicate the approach of a train. Approach Lighting: A method of lighting signals upon the approach of a train. (See also Continuous Lighting Control). Approach Locking: (See Electric Locking). Approach Signal: (See Signal). Approach Track: (See Route Signaling). ARA: (Historical) American Railway Association. AREA: American Railway Engineering Association. AREMA: American Railway Engineering and Maintenance of Way Association North American body for determination of railway engineering standards, equivalent to the European International Union of Railways (UIC).

18-6

Glossary

Armature: A piece of steel, soft iron, or a coil, so placed as to be acted upon by an electromagnet or a permanent magnet; or that part of an electric generator in which electricity is generated; or that part of a signal motor which rotates. Arrester: (See Surge Protective Device) ASA: (See ANSI). ASC: Automatic Speed Control. ASCE: American Society of Civil Engineers. ASME: American Society of Mechanical Engineers. Aspect: (Signal Aspect) The signaling term meaning the indication given to the driver of an approaching train by a signal as in the signal was showing a green aspect. ASTM: American Society for Testing & Materials. Asynchronous Motor: Modern traction motor type using three phase electrical supply and now the favored design for modern train traction systems. Can be used on ac and dc electrified railways with suitable control electronics and on diesel-electric locomotives. ATC: Automatic Train Control. ATO: Automatic Train Operation. ATP: Automatic Train Protection. Automatic Block Signal System: A series of consecutive blocks governed by block signals, cab signals, or both, actuated by a train, or engine, or by certain conditions affecting the use of a block. Normal Clear: A term used to express the normal indication of the signals in an automatic block system in which an indication to proceed is displayed except when the block is occupied. Normal Stop: A term used to express the normal indication of the signals in an automatic block system in which the indication to proceed is given only upon the approach of a train to an unoccupied block. (See also Absolute Permissive Block System, Centralized Traffic Control, Overlap Block Signal System and Traffic Control System). Automatic Signaling: A system whereby lineside visual signals for train drivers are controlled entirely by the passage of trains using track circuits Used where there are no junctions or turnouts, which require semi-automatic signaling.

18-7

Glossary

Automatic Switching: A term generally used to describe a system of controls for automatic operation of track switches whereby routes that are established by preliminary manual selection are automatically completed by the progress of cars or trains. (See also Route Selection). Auxiliary Supply: System for providing power for vehicle or train lighting, air conditioning, heating, on-board catering or passenger facilities and emergency battery systems. Generally a train has two power supply systems, one for traction and the other for auxiliaries. Auxiliary systems can be supplied from the main traction power source through a generator or converter, or through an axle driven generator. AWG: American Wire Gage. AWS: Automatic Warning System. Axle: The circular shaft connecting two wheels to form a wheelset. The wheels are an interference fit to ensure the gauge is maintained. Wheels are removed by forcing them off after injecting oil under high pressure into the wheel hub through a specially designed aperture drilled in the hub. Axlebox: The housing attaching the axle end to the bogie that contains the bearing allowing the axle to rotate. Axle Brush: Axle mounted contact used to provide a return path for power circuits on electric vehicles. Nowadays normally mounted on the axle end as part of the axlebox assembly. Axle Counter: An automatic arrangement for detecting and counting car and locomotive axles that pass a given wayside location: usually makes use of a wheel detector. B-H Curve: (See Magnetization Curve and Hysteresis). Back Contact: (See Contact). Back Light: (Highway Grade Crossing Signal) An auxiliary signal light used for indication in a direction opposite to that provided by the main unit. Back Light: (Signal) The light from a signal electric lamp, visible through a small opening in the back of a light signal, used for checking the operation of the signal lamp. Ballast: (Track) The material most commonly used to form the roadbed of a railway track. It is laid on the base formation of the track with the track laid on top of it and provides a storm water drainage medium. It usually consists of granite, whinstone or furnace slag. Ash is sometimes used in yards but not where any sort of speed is required as the dust gets into the rolling stock equipment. A properly maintained railway will have regular tamping of ballast to ensure the track itself is maintained to provide an acceptable standard of ride. Other track forms include slab or non-ballasted track that does not require ballast.

18-8

Glossary

Ballast-Lighting: (Power) An electrical device used to assure the proper operation of fluorescent lamps by stabilizing the arc. Traditional ballasts are reactances that make use of conventional magnetic components. New electronic ballasts make use of solid-state electronic elements that allow higher-frequency operation of the lamps and significantly reduce power consumption. Balanced Currents: (Power) A set of conductors whose currents instantaneously sum to zero are said to have balanced currents. A three-phase electric power transmission line is said to have balanced currents if the phase currents are equal in magnitude and at 120 electrical degrees with respect to each other (only positive-sequence symmetric components). Ballast Resistance: (Railroad) The resistance offered by the ballast, ties, etc., to the flow of leakage current from one rail of a track circuit to the other. (See also Leakage Current). Banjo: Common UK name for the drivers vacuum brake control handle, arising from its shape. The term is also sometimes used to refer to a disc shaped ground signal. Bank: (UK) A section of railway line on a gradient. In North America this is known as a grade. Banner: (Historical) The actuated part of a disc or wig-wag signal. Bare Illuminant: (Historical) The light produced by a kerosene flame or electric lamp. (See also White Light). Battery: (Railroad) Locomotives and carriages are nowadays invariably fitted with large, heavy-duty batteries to provide start up and emergency power. Batteries are usually recharged from a power line running along the train or from the traction power supply through a generator, alternator or converter. Battery: (Voltage) (Historical- Railroad). High Voltage: Usually refers to a 110-volt battery. Low-Voltage: Usually refers to a battery of 30 volts or less. Battery Charge: (See Charge). Battery Charger: A device used to replenish a battery by changing alternating current to direct current that matches the voltage of the battery. Battery Chute: (Historical) A small cylindrical receptacle for housing track batteries and set in the ground below the frost line. Battery Elevator: (Historical) An arrangement of shelves in a supporting frame by means of which batteries may be lowered into, held in position, and raised out of a battery chute. (See also: Battery Chute). Battery Locomotive: Some electric railways, particularly those with long underground sections like the London Underground or the Hong Kong Mass Transit Railway, employ battery powered locomotives for use on engineers trains at night when the main traction power supply is 18-9

Glossary

switched off. They are arranged to recharge from the traction power supply during the daytime when they are not required for engineers trains. Their disadvantage is the time required out of service for recharging and the additional weight of the batteries. Battery Well: (Historical) (Railroad) A housing, usually of concrete, set in the ground for housing batteries below the frost line. Bed: (See Locking Bed) (Historical). Bell: (Highway Grade Crossing) An audible device at a highway grade crossing that provides a warning. Bell Code: (Historical) A code in which the strokes of a bell have a predetermined significance. Bi-Directional Signaling: (UK) Used to allow trains to run in either direction over the same section of track subject to the built-in safety systems that prevent collisions. Bi-directional signaling is very useful in releasing for maintenance a single track of a two-track railway but it is more complex and expensive to install than single direction signaling. Binary Alloys: A number of commercial alloy materials consisting primarily of two components (and sometimes the addition of minor constituents) have been used where high permeability is required. These binary alloys include iron-silicon, iron-nickel, and iron-cobalt. Biot-Savart Law: Ampres law can be used to calculate magnetic fields (or flux density) only if the symmetry of the current distribution (current carrying conductor) is high enough to permit the easy evaluation of the line integral B dl . This requirement limits the usefulness of the law in many practical problems. The law does not fail; it simply becomes difficult to apply in a useful way. Another approach to computing the magnetic field at a point for an arbitrary charge distribution is to divide the charge distribution into current elements. The law of Biot and Savart is then used to calculate the field (flux density) contribution, dB, due to each current element at the point in question. The law of Biot and Savart may be written in vector form as:
dB =

i dl r 4 r 3

The resultant field at a point is then found by integrating the field (flux density) contributions for the entire charge distribution as: B = dB. The law of Biot and Savart is used to compute the magnetic field at a point for a system of conductors with complex geometry (e.g. the catenaryshaped conductors of a high voltage transmission line, substation buswork, etc.) This law is sometimes attributed to Biot and Savart, sometimes to Ampre, and sometimes to Laplace, but no one of them gave it in its differential form. (See also Ampres Law). Blade: The moving part of a point or rail switch that causes a train to change direction at a diverging junction. Blade-Grip: (Historical) That part of a semaphore arm or spectacle to which the blade is secured. Block: A length of track of defined limits, the use of which by trains and engines is governed by block signals, cab signals, or both. (See also Station). 18-10

Glossary

Absolute: A block in which no train is permitted to enter while it is occupied by another train. Permissive: A block in manual or controlled manual territory, based on the principle that a train other than a passenger train may be permitted to follow a train other than a passenger train in the block. Block (UK): In signaling terminology, a physical length of track protected by a fixed signal which indicates to a driver when it is safe to proceed into the section. Block Indicator: An indicator used to indicate the condition of a block. (See also Switch Indicator, Track Indicator). Blocking: A means to prevent display of a signal when it is desired to inhibit entry of a train movement into the block governed by the signal. (See also Field Blocking, Office Blocking, Track Blocking). Blocking Back (UK): A term used to denote a queue of trains detained behind a delayed train or other obstruction. Block Signal: (See Signal). Block Signal System: A method of governing the movement of trains into or within one or more blocks by block signals or cab signals, or both. (See also Automatic Block Signal System, Manual Block Signal System). Block System: A series of consecutive blocks. Blocking Device: (See Lever Blocking Device). Bogie (UK): A 4- or 6-wheeled truck used in pairs under long-bodied railway vehicles. The bogie has a central pivot point that allows it to turn as the track curves and thus guide the vehicle into the curve. In North America it is always referred to as a truck. There are almost as many bogie designs as there are bogies. Bond: A low impedance connection to assure the required electrical conductivity between conductive parts to be electrically connected. Bonding: The electrical interconnecting of conductive parts designed to maintain a common electrical potential. Bonding Connection: A reliable conductor used to make a connection to ensure the proper electrical conductivity between metal parts that are required to be electrically connected. Bonding Screw: A screw used to connect (bond) the neutral bus of an electrical panel to the panels grounded case. If the electrical panel is located at the service entrance the screw should 18-11

Glossary

be turned in, so that the neutral bus is grounded. In a sub-panel, however, the screw should be turned out because the neutral bus should not be grounded at the sub-panel per the NEC. Bolt Lock: (Railroad) A mechanical lock so arranged that if a switch, derail or movable point frog is not in the proper position for a train movement, the signal governing that movement cannot display an aspect to proceed: and that will prevent a movement of the switch, derail or movable point frog unless the signal displays its most restrictive aspect. Bootleg: (Railroad) A protection for track wires where the wires leave the conduit or ground near the rail. Boundary Element Methods: The methods to calculate electric and magnetic fields based on simulating the presence of boundaries between different materials with sets of charges or current distributions that become the unknown quantities to be calculated. For instance, the perturbation of the electric field caused by a conductive object may be simulated by a set of charges placed on or below the surface of the object. The distribution and value of these charges are determined by calculating the known potential on various points of the conductor surface. Similarly, the perturbation of the magnetic field caused by a ferromagnetic object may be simulated by a set of dipole moments. Box (UK): Railway term for signal box or signal cabin. Known as a Tower in the U.S. Boxing (UK): Railway term for the sideways motion of a bogie (truck on North American railroads) that has a damaging effect on rails. Also known as hunting. Bracket Signal: (See Signal, Types and Arrangements). Brake Application: (See also Dynamic Braking). Full Service: An application of the brakes resulting from a continuous or a split reduction in brake pipe pressure at a service rate until maximum brake cylinder pressure is developed. As applied to an automatic or electro-pneumatic brake with speed governor control, an application other than emergency that develops the maximum brake cylinder pressure, as determined by the design of the brake equipment for the speed at which the train is operating. Braking Distance: The maximum distance on any portion of any railroad which any train operating on such portion of railroad at its maximum authorized speed, will travel during a full service application of the brakes, between the point where such application is initiated and the point where the train comes to a stop. Branch Circuit: (Power) The circuit conductors in building wiring between the final overcurrent device protecting the circuit and the outlets served the circuit. Breakaway (UK): Term for the accidental breakage of couplings that results in a train becoming divided. The brake system is arranged to ensure that both portions of the train will stop in such an event.

18-12

Glossary

Breakdown Voltage, AC (Peak) (Railroad) The minimum peak value of a sinusoidal voltage at frequencies between 15 Hz and 105 Hz that results in diversion of current through the surge protective device. Breakdown Voltage, AC (RMS): (Railroad) The minimum rms value of a sinusoidal voltage at frequencies between 15 Hz and 105 Hz that results in diversion of current through the surge protective device. Breakdown Voltage, DC: (Railroad) The minimum slowly rising dc voltage that results in diversion of current through the surge protective device. Bungalow: (See Housing). Bus: A conductor(s) that serves as a common connection for the corresponding conductors of two or more circuits. In an electric power substation, it is a connection between components of electrical substations such as switchgear, transformers, and exit lines. In the power substation, a bus is generally three-phase and is composed of flexible or rigid conductor sections mounted on insulators. Cab: The compartment of a locomotive from which the propelling power and power brakes of the train are manually controlled. Cabin: (Historical) A common name for an interlocking or block station. Cab Indicator: (Railroad). Audible: A device located in the cab designed to sound under predetermined conditions. Visual: A signal located in the cab indicating a condition affecting the movement of a train or engine. Cab Signal: (See Signal). Cab Signaling (UK): A system where signal indications are displayed to the driver in his cab, usually from a trackside induction system. Cab signals may either augment existing lineside signals or replace them. Cab signaling is usually used in conjunction with some sort of automatic protection system that prevents a driver allowing his train to enter an occupied section or exceeding a predetermined safe speed. Cab Signal System: A system that provides for the automatic operation of cab signals. Cable Jacket: A protective covering over the insulation, core, or sheath of a cable. Cable Post: (Railroad) An upright designed for supporting a cable. Cable Sheath: A conductive protective covering applied to cables. Caboose: A North American term for brake van. 18-13

Glossary

CAD: (See Computer Aided Dispatching). Calling-On Signal (UK): Used to indicate to a driver that he may proceed at caution speed because he is entering a section occupied by another vehicle or train. Normally used for coupling operations at station platforms. Calrod: (Railroad) An electric switch heater. (See Switch Heater). Canadian Electrical Code Part Three (CEC): A series of standards governing the use of electric supply and communications lines, equipment and associated work practices employed by an electric supply, communications, railway, or similar electric company in the exercise of its function as an electric company. It is sponsored by the Canadian Standards Association (CSA C.3). Cancellation Loops: Cancellation loops is the general name given to the use of conductive wires arranged in paths (loops) near to transmission or distribution lines (or other linear sources) to reduce magnetic fields within a region of interest. The magnetic field due to the power line induces currents in the conductive loops that, in turn, create another magnetic field that partially cancels the original magnetic field. The magnetic fields due to induced currents in the cancellation loops are in the same direction but with a phase difference with respect to the source currents. Sometimes capacitance can be added to the loop to enhance the cancellation effect by inducing larger loop currents to flow that generate larger cancellation fields. The geometry of the loop will depend on the power line or other source geometry. Although the cost of cancellation loops is small compared to the cost of a power line, cancellation loops are expensive to install, they have losses associated with them, magnetic field is reduced in a region but could be increased in another region, and they must be designed and maintained in accordance with the NESC and all applicable safety codes. There are two types of cancellation loops: active and passive loops. Passive cancellation loop currents are induced by the source to be shielded, whereas active loop current is supplied by an independent source and may use a feedback control system to adjust loop current to match magnetic field changes within the region to be shielded due to changes in source current. Cantilever: (Railroad) A structure, consisting of a ground mast and a horizontal arm extending to one side, used to support one or more signals as required for multiple tracks, or one or more highway grade crossing signals. Capacitor: A device which consists essentially of two conductors (such as parallel metal plates) insulated from each other by a dielectric and which introduces capacitance into a circuit, stores electrical energy, blocks the flow of direct current, and permits the flow of alternating current to a degree dependent on capacitance and the current frequency. Capacitance: The property of a system of conductors to store electric charges. It is expressed as the ratio between an electric charge and a voltage difference (C/V). Its unit is the farad (F). (See also Farad). Capacitor Bank: Groups of capacitors on the distribution and transmission systems. Typically use to provide voltage support or VAR compensation. All three phases have matching capacitors wired phase-to-ground. Capping: (Railroad) (Historical) (See Trunking) The covering for trunking. 18-14

Glossary

Case: (See Housing). Catch Up (Railroad- Hump Yard) A car or cut of cars descending the grade from the apex of the hump and overtaking a preceding car or cut of cars before it has cleared the detector track circuit of the last switch in the group. Catenary: (Power) The curve assumed by a flexible conductor or other wire hanging freely between two supports and loaded uniformly throughout its length. Catenary: (Railroad) Originally the term used to denote an overhead power line support wire, derived from the curve a suspended wire naturally assumes under the force of gravity. Now adopted to mean the whole overhead line system. See overhead. Central Instrument Location: (See Housing). Centralized Traffic Control: A term applied to a system of railroad operation by means of which the movement of trains over routes and through blocks on a designated section of track or tracks is directed by signals controlled from a designated point without requiring the use of train orders and without the superiority of trains. (See also Traffic Control System). Centre Siding (UK): A track laid between two running lines for the purpose of reversing trains, usually beyond a station. It allows a train to reverse direction without crossing a track carrying through trains. Sometimes referred to as a reversing siding. In North America it is referred to as a pocket track or a turnback track. Charge: (Power) (See Electric Charge). Charge: (Railroad) Battery: The restoration of the active materials in a battery by passing a unidirectional current through it in the opposite direction to that of the discharge. Battery Boost: A partial charge, usually at a high rate for a short period. Constant-Current: A charge in which the current is maintained at a constant value. Constant-Voltage: A charge in which the voltage at the terminals of the battery is held at a constant value. A modified constant-voltage charge is the one in which a fixed resistance is installed in the charging circuit to limit the initial current. Equalizing: An extended charge that is given to a storage battery to insure the restoration of the active material in all the cells. Floating: Maintaining a storage battery in operating condition by a continuous charge at a low rate or primary battery coupled with a rectifier to provide a supplementary source of energy. 18-15

Glossary

Initial: A term used in expressing the first charge given a storage battery after it has been set up, also to designate that recommended current applied to electrical apparatus at the beginning of a series of tests. Taper: A charge in which the charging current tapers from a high to a low value as the charge progresses and the voltage of the battery rises. Trickle: A continuous input of current to a storage battery to compensate for internal losses only. Charge Density: In space (space charge density) is the electric charge per unit volume (C/m ). On a surface (surface charge density) is the electric charge per unit area (C/m2). The electric field on a conductive surface with charge density is given by E = /. ( is the permittivity of the dielectric medium). Charge Simulation Method: A method used to calculate electric fields in which charges are placed near or at the surface of electrodes at strategic locations to simulate the effect of the electrodes. Equations are written that relate the voltage at a number of points on the surface of the electrodes, the potential points, to the charges. If the system can be described with N charges and N potential points are selected, the solution of the N equations provides the value of the charges. Electric fields and space potentials are calculated from the charges so determined. Chart: (See Track Indicator Chart) (Historical). Choke Coil: (Railroad) A form of stationary induction apparatus to supply reactance and used commonly in connection with signal lightning arresters. Chopper Control (UK): A development in electric traction control which eliminates the need for power resistors by causing the voltage to the traction motors to be switched on and off (chopped) very rapidly during acceleration. It is accomplished by the use of thyristors and will give up to 20% improvement in efficiency. Chute: (See Battery Chute) (Historical). Circuit Controller: A device for opening and closing electric circuits. Movable Bridge: A device for opening and closing electric circuits between the stationary and movable bridge spans. (See Also Movable Bridge Coupler). Switch: A device for opening and closing electric circuits operated by a rod connected to a switch, derail or movable point frog. (See also Point Detector). Circular MIL: An area equal to that of a circle with a diameter of 0.001 inch (one mil). It has been used to specify the cross section of conductors. Circulation: The mathematical quantity, circulation, is the line-integral of a vector field along a given path. 18-16
3

Glossary

Clamp-On Ammeter: An instrument to measure the alternating current in a conductor. The instrument consists of a magnetic circuit that can be opened and clamped on the conductor and of a voltmeter that measures the voltage induced in a wire wound on the magnetic circuit. Clamping Voltage: (Railroad) The maximum magnitude of voltage across a surge protective device (SPD) during the passage of a specified surge current under specified test conditions, e.g. dynamic, static, phase angle, pulse polarity. Clear: (Railroad) Free from obstruction or other restricting feature. Clear Block or Track: Unoccupied section of track. Clear Signal: The name of a signal aspect, the indication of which is defined in the operating instruction of the railroad. Clearance Point: (Railroad) The location on a turnout at which the carriers specified clearance is provided between tracks. (See also Fouling Point). Clearance Track Circuit: (See Track Circuit). Closed Circuit Principle: The principle of circuit design where a normally energized electric circuit which, on being interrupted or de-energized, will cause the controlled function to assume its most restrictive condition. Code: (Railroad) The controlled pulsing of electrical energy in a line or track circuit, usually for the purpose of transmitting information. The pulses may be on/off or polarized, or both, and may also vary in duration. (See also Track Circuit, Standard Code). Code Converter: (Railroad) A group of relays and circuits used in conjunction with field carrier apparatus for conversion and coordination between a carrier circuit and a remotely controlled direct current line. Code Follower: (See Relay- Code Following) (See also Code Repeater). Code Line: (Railroad) A non-vital line circuit, the principle purpose of which is to carry control and/or indication codes, such as for the supervisory control system portion of Centralized Traffic Control (CTC). Code Repeater: An arrangement for retransmitting received coded electrical energy. Code Transmitter: A device to vary periodically an electrical circuit at a definite predetermined code frequency. Code System: The non-vital apparatus and circuits used for forming, transmitting, receiving and applying the codes of a supervisory control system. (See also Application Unit). Coercive Force: In a typical magnetization curve (or hysteresis loop), if the magnetic field, H, applied to a specimen is increased to saturation and is then decreased, the flux density B decreases, but not as rapidly as it increased along the initial-magnetization curve. Thus, when H reaches zero, there is a residual (remanence) flux density (Br) or offset from the origin. To reduce B to zero, a negative field (-Hc) must be applied- this field is called the coercive force. If the field, H, originally applied to the specimen is high enough to drive the specimen to saturation, the 18-17

Glossary

resulting coercive force is called the coercivity. The coercivity of material is a characteristic of the material (such as resistivity). (See also Hysteresis, Magnetization Curve and Remanence). Coercivity: (See Coercive Force and Hysteresis). Common-Mode Voltage: (Longitudinal) The voltage that appears equally in amplitude and in phase from each signal conductor to ground or to a common reference. Common Return Circuit: A term applied where one wire is used for the return of more than one electric circuit. Compact Line: A power line for which the distances between phases are much less than those used in conventional designs. This is made possible either using special insulators or reducing the overvoltages applied between phases or reducing the flashover reliability of the line. Compact Multipole: A power line that can be considered a two-dimensional multipole and uses the minimum possible number of conductors. Examples of compact multipoles are given in the figure below.

B = Magnetic Field - mG I = Current - A

P = Phase Spacing - m R = Distance - m

Example of Compact Multipoles Computer Aided Dispatching: A term relating to use of computers in centralized traffic control systems to aid the dispatching of trains. Condenser: (Historical) (See capacitor). Conductivity: Conduction current occurs in the presence of an electric field within a conductor. The conductivity () of a material reflects the relative ease with which the current moves through the conductor. Conductivity is the reciprocal of resistivity () and both are characteristics of a material rather than a particular specimen; it is defined for isotropic materials in units of siemens per 18-18

Glossary

meter (S/m) and sometimes in mhos per meter ( /m); like resistance, conductivity is a function of temperature. Conductor: (Power) A conductor is a material within which charge is free to move. Metals and electrolytes are conductors; the flow of charge within a metal or electrolyte is governed by Ohms law at a point: current density is equal to the product of material conductivity and electric field in the material (See equation under Current Density). Conductor: (Railroad) North American term for train staff member whose duties vary with different administrations but who is primarily responsible for train and passenger safety. Often also involved with ticket inspection. In the UK this term was often referred to as the guard but the same term is now common. Conductor Rail (UK): (Railroad) An additional rail (or rails) provided on those electric railways where power is transmitted to trains from the track. Often referred to as the third rail or current rail, it is normally at positive potential and is mounted on insulators to the outside of and slightly higher than the running rails. The return of the circuit is via the running rails. The current is collected by the train through shoes, attached to the bogies, which slide along or under the rail. Varieties of the system include top, side and bottom contact rails. Top contact rails are susceptible to ice and snow contamination in cold climates and present a certain risk to persons walking on the track. For this reason, bottom contact rails are preferred for modern systems. The continuity of conductor rails must be broken at junctions in the track to allow continuity of the running rails. Such gaps may cause momentary loss of power to the train. There are cases from time to time of trains becoming gapped at complex junctions, i.e. they stall over a gap and have to be rescued by another train. London Underground has a fourth rail (negative) for a completely insulated circuit. It is known as a four-rail system. Conflicting Movements: (Railroad) Movements over conflicting routes. Conflicting Routes: Two or more routes, opposing, converging, or intersecting, over which movements cannot be made simultaneously without possibility of collision. Consist: North American term for train formation e.g. This vehicle was in the consist A consist is a listing showing the train number, the dates and times of departure as well as arrival; the locomotive, radio and caboose number; the initial and car numbers of each car on the train; the billing of these cars; the special handling of the cars and the name of the conductor. It reflects all activities that took place on the movement of cars between any two stations. Console: (Railroad) An assemblage of manually operated levers or other devices for the control of signals, switches or other units, without mechanical interlocking, usually including a track diagram with indication lights. (See also Machine). Constant Warning Time Device: (Railroad) A device used as a part of a highway grade crossing warning system to provide a relatively uniform warning time. (See also Motion Sensitive Device). Contact: (Railroad) A conducting part that co-acts with another conducting part to open or close an electric circuit. 18-19

Glossary

Back: A part of a relay against which, when the relay is de-energized, the current carrying portion of the movable neutral member rests so as to form a continuous path for current. Closed: A current-carrying member that is closed when the operating unit is in the normal position. Dependent: A contacting member designed to complete any one of two or three circuits, depending on whether a two or three-,way device is considered. Front: A part of a relay against which, when the relay is energized, the current-carrying portion of the movable neutral member is held so as to form a continuous path for current. Heel: (See Dependent). Independent: A contacting member designed to complete one circuit only. Magnetic Blow Out: A contact which is fitted with a special device, such as a permanent magnet, to aid the interruption of any arc that is drawn upon contact opening. Normal: A term used to designate a current-carrying member when the operated unit is in the normal position. Open: A current-carrying member that is open when the operating unit is in the normal position. Polar: A part of a relay against which the current-carrying portion of the movable polar member is held so as to form a continuous path for current. Reverse: A term used to designate a current-carrying member when the operated unit is in the reverse position. Contact Resistance: The resistance produced by the contact of two surfaces. Continuous Control: (Cab Signal, Train Control, etc.) A type of control in which the locomotive apparatus is constantly in operative relation with the track elements and is immediately responsive to a change of conditions in the controlling section which affects train movement. (See also Intermittent Control). Continuous Lighting Control: A method of lighting signals continuously. (See also Approach Lighting). Control: (Railroad) (See Centralized Traffic Control, Console, Continuous Control, Dual Control, Intermittent Control, Remote Control, Semi-Automatic Control, Station, Traffic Control System, Train Control System) Control Circuit: An electrical circuit between a source of electric energy and a device that it operates. Control Operator: (Railroad) An employee assigned to operate the control machine of a traffic control system.

18-20

Glossary

Controlled Point: (Railroad) A location where signals or other functions or both of a traffic control system are controlled from the control machine. Controlled Siding: (See Siding). Controlling Section: (Historical) One or more track circuit sections governing approach to or movement within a block. Converter: Electronic system for converting alternating current to direct current. Used, for example, where an ac supply has to be converted to dc for battery charging purposes. Also was used for converting ac traction supplies required for dc traction motors. (See also Inverter.) Corona: (overhead power line corona and radio noise) A luminous discharge due to ionization of the air surrounding an electrode caused by a voltage gradient exceeding a certain critical value. Electrodes may be conductors, hardware, accessories, or insulators. Coulomb: A unit of electric charge named after the French physicist Charles Coulomb (1736-1806). A coulomb is equal to the quantity of electric charge carried by one ampere of 18 current in one second. A coulomb of charge is equal to the charge of 6.25 x 10 electrons. Coulombs Law: Coulombs Law states that the force between two electric charges (or a pair of magnetic poles) is directly proportional to the product of their magnitudes and inversely proportional to the square of the distance between them. Counterpoise: (Power) A system of buried conductors around the base of transmission line structures that is connected to the grounding system of the towers or poles supporting the line. Its function is to dissipate a lightning discharge to ground over a larger area, thereby lessening the resistance to the flow of these discharges. It is also used sometimes to solve EMC problems by providing passive shielding from magnetic field induction. Coupling: (Electrical) Energy transfer among circuits, equipment or systems caused by electromagnetic interaction. There are three forms of coupling: magnetic (inductive), electric (capacitive), and conductive (resistive). In addition, coupling by electromagnetic radiation exists and is associated with propagation of radiation fields, e.g., radio frequency interference (RFI). Covered Conductor: A conductor covered with a dielectric (insulating material) having no rated insulating strength or having a rated insulating strength less than the voltage of the circuit in which the conductor is used. Crest Factor: The ratio of a waveform crest (i.e. peak, or maximum) value to its root-mean-square (rms) value. For example, for a purely sinusoidal waveform the crest factor is: Bmax/Brms = 2 Highly distorted or pulsed waveforms can have much larger crest factors. If shielding is designed for average or rms values (based on measurements), a significantly distorted waveform could 18-21

Glossary

saturate the shield if the crest factor is not properly considered in the shield design. Some typical crest factors are (Analog Devices 1992 Special Linear Ref. Manual):
Sine Wave: Square Wave: Triangular Wave: Gaussian Noise: Rectangular Pulse Train: SCR Waveform: 50% Duty Cycle: 25% Duty Cycle: 2-10 4.7 1.414 1.0 1.73 3.0 2-10, or more

Cripple (UK): Defective vehicle or train. This term is known in the U.S. as a Bad Order. Cross-ARM: (Railroad) An arm, usually fastened at right angles to a pole, designed to carry the pins and insulators to which wires may be attached. Crossover (UK): A track providing a connection between two parallel tracks using two sets of points. A scissors crossover provides two connections, one in each direction. Cross-Phasing: (See Low Reactance Phasing). Cross Protection: (Railroad) An arrangement to prevent the improper operation of a signal, switch, movable point frog, or derail as the result of a cross in electrical circuits. Crossing Signs: (See Highway Grade Crossing Sign). (Also referred to as Crossbucks). Crossing Warning Device: (See Highway Grade Crossing Warning Device, Highway Grade Crossing Signal). Crossover: (Railroad) Two turnouts with the track between the frogs arranged to form a continuous passage between two nearby and generally parallel tracks. CTC: Centralized Traffic Control. CTC (Canada): Canadian Transport Commission. Curie Temperature: When the temperature of a ferromagnetic material is raised above a certain critical value, called the Curie Temperature (or Curie Point), the materials magnetic properties change and the material becomes merely paramagnetic. The change is brought about when the charge coupling between adjacent atoms (causing domain regions) suddenly disappears. The Curie temperature is usually lower than the melting point of the substance. This means that ferromagnetic material that has been welded will lose its high permeability in regions exposed to the high heat. The following are some examples of Curie temperatures: Iron = 770C, cobalt = 18-22

Glossary

1131C, nickel = 358C, gadolinium = 16C). (See also Ferromagnetism, Paramagnetism, and Permeability). Current Density: If an electric current (I) is distributed uniformly across a conductor of cross-sectional area (A), the magnitude of the current density vector (J) for all points on that cross section is: J = I/A. The term refers to either to conduction-current density or to displacement-current density, or to both. The MKS (SI) system unit of J is ampere per square 2 meter (A/m ). Current density is proportional to the conductivity () and electric field (E) in a medium (Ohms Law at a point) such that: J = E. If the current is confined to the surface, a surface current density is defined. Current of Traffic: The movement of trains on a main track, in one direction, as specified by the rules or in the timetable. Current Rail (UK): See conductor rail. Cut: North American term for uncoupling. Cut was also used in the UK to refer to the sections a train was broken into when passing over a hump in a marshalling yard. Since there isnt any hump shunting in the UK any more this term may be obsolete. Cut-In Circuit: (Railroad) A roadway circuit at the entrance to automatic train stop, train control or cab signal territory by means of which locomotive equipment of the continuous inductive type is actuated so as to be in operative condition. Cut-Out: (Railroad) The means provided in signal system control logic to temporarily suspend an indication that will be reinitiated at some point in the operation of the system. Cut-Section: (Railroad) A location other than a signal location where two adjoining track circuits end within a block. Relayed: A cut-section at which the energy supply to one track circuit is controlled by the adjoining track circuit. Dark Territory: (Railroad) Non-signaled trackage or territory. This means any part of a rail system that does not have railroad tracks equipped with signal equipment to safely direct the movement of trains. The determination of signaled vs. non-signaled depends only on the presence of a signal system for the trains and does not depend on the presence or absence of grade crossing warning systems, or other signal equipment. (See also Signaled Territory). DC (Direct Current): The term dc is applied to both currents and voltages that are non-alternating and change little with time. They may be positive or negative. Dead Section: A section of track, either within a track circuit or between two track circuits, the rails of which are not part of a track circuit. Decoder: A device adapted to control apparatus in a manner corresponding to the code to which the track or line circuit is subjected. (See also Decoding Transformers). Decibel: A numerical expression of the relative differences between some quantity and a reference level. For a quantity of power (e.g. watts or sound pressure) the decibel (dB) level is equal to ten times the common logarithm of the ratio of a level of interest to a reference level. 18-23

Glossary

For linear quantities (e.g. volts, volts/meter, amps/meter, etc.), the decibel level is twenty times the common logarithm of the ratio of a level to a reference level. De-Energized: A conductor, current carrying component, or other object that is free from any electrical connection to a source of potential difference and from electric charge so that it does not have a potential different from that of the earth. De-Energized Position: The position assumed by the moving member of an electromagnetic device when the device is deprived of its operating current. Defect Detector: (Railroad). Broken Flange: A device capable of detecting a missing section of the flange area of a wheel. Dragging Equipment: A device capable of detecting equipment dragging from a passing train. Deflecting: (See Roundel). Degaussing: Shielding material can sometimes experience a reduction in permeability because the residual induction, or retentivity, may capture the domains in the material. Degaussing restores the captured domains by exposing the material to a very large ac field and slowly decreasing the amplitude to zero. To demagnetize a magnetic material completely, the residual induction (retentivity) must be reduced to zero. This usually cannot be accomplished by a reversed dc magnetizing force, because the material then would just become magnetized with opposite polarity. The practical way is to magnetize and demagnetize the material with a continuously decreasing hysteresis loop. This can be done with a magnetic field produced by alternating current. Then as the magnetic field is continually reduced, the hysteresis loop becomes smaller and smaller. Finally, with the weakest field, the loop collapses practically to zero, resulting in zero retentivity (residual induction). This method of degaussing is also called demagnetization. One application is degaussing the metal electrodes in an older color picture tube with a degaussing coil using alternating current from a typical electrical outlet. Another example is erasing the recorded signal on magnetic tape by demagnetizing with an ac bias current. (See also Domain Capture Effect, Hysteresis Loop, and Magnetization Curve). Delay Time: The time that elapses after an automatic brake application is initiated until the brakes start to apply. Delta Configuration: (Power) An electric power transmission line conductor configuration in which the three electrical phase conductors are physically arranged in a triangle or delta shape. In a horizontal delta configuration, the triangle apex is aligned in a horizontal or sideways direction; whereas in a vertical delta configuration, the triangle apex points either up or down in the vertical direction. If the middle phase is lower than the other two, the configuration is called inverted delta or, by some, nabla.

18-24

Glossary

Delta Connection: (Power) A delta connection is a method of connecting the windings of a transformer in series so that a closed circuit is formed. Delta windings prevent the flow of zero sequence currents. Demagnetization: (See Degaussing). Detection: In railway signaling, the ability to determine that a train occupies a track section or block. Detection is usually by a track circuit or equivalent electrical loop. Also used to verify that a point or signal has operated correctly as part of the interlocking. Detector: (See also Defect Detector). Dragging Equipment: A device used to equipment dragging along the tracks. High Water: A device used to detect flooding on or near the tracks. High-Wide Load: A device capable of detecting loads of excessive height or width on a passing train with respect to accepted track structure clearances. Hot Wheel: A device capable of detecting abnormal wheel temperatures on passing trains, generally caused by sticking brakes or improperly set hand brakes. Loose Wheel: A device capable of detecting excessive lateral wheel movement on passing trains. Motion: (See Motion Sensitive Device). Overheated Journal: A device capable of detecting abnormal heating in wheel bearings (called a hot box) on passing trains. Presence: A device that detects the presence of an axle, wheelset, or train on the track. Slide: A device used to detect a landslide on or near the tracks (also called a slide fence). Wheel Defect: A device used to measure the degree of impact caused by a flat spot or other wheel defect. Detector Track Circuit: (See Track Circuit). Detonators: (UK) Explosive devices carried on locomotives or trains for use in emergency situations where they are placed on the top of the running rail so that they explode to give a warning to the driver if a train runs over them. First appeared in 1841. Known in North America as torpedoes. Diamagnetism: Most materials show some magnetic effects due to electric charges in motion at the molecular, atomic, and subatomic levels. (e.g. electrons revolve around the nucleus of an atom and also spin about their own axis as they move as paired or unpaired electrons). With the exception 18-25

Glossary

of the ferromagnetic group of materials, these magnetic effects are weak. Depending on their magnetic behavior, materials can be classified as diamagnetic, paramagnetic, or ferromagnetic. In diamagnetic materials, a magnetic moment is induced into the materials atomic structure by an applied magnetic field in a direction opposite to that of the applied field. Zinc, bismuth, sodium chloride, lead, water, gold, and mercury are a few of the materials that are feebly repelled by a strong applied magnetic field, and are said to be diamagnetic. This type of behavior is called diamagnetism. Diamond: A place where two railway tracks cross each other. Diamond Crossing: The point where two tracks cross without connection. It is named after the shape of the track formation that occurs. Dielectric: A term often used to denote an insulating material. A dielectric is so called because it permits the passage of the lines of force of an electric field (electric flux) but does not conduct an electric current. Dielectric materials can therefore store electric charge. (See also Permittivity). Dielectric Breakdown Voltage: The voltage that will result in breakdown or sparkover when impressed across a dielectric. Dielectric Constant: A term used to describe a materials ability (with respect to a vacuum) to store electrostatic energy per unit volume for a unit potential gradient. (See also Permittivity). Dielectric Union: A non-conductive connection between two parts of a conductive pipe. A dielectric union maintains the function of the pipe but interrupts its electrical continuity. Dielectric Withstand Level: The ability of insulating materials; spacing or equipment to withstand a specified voltage for a specified time without breakdown or sparkover. Diesel Locomotive: A locomotive whose principle power source is a diesel engine. This engine drives the wheels either by mechanical transmission or electrical transmission. Electrical transmission is by far the most common. Diesel locomotives with electrical transmission are known as Diesel-Electric locomotives. The diesel engine drives a generator (or an alternator on more modern machines) that produces the current required to power the electric traction motors that drive the axles. Differential Mode Voltage (Transverse): The voltage at a given location between two conductors of a group, not between ends of the same conductor. Digital Multimeter (DMM): A piece of test equipment used for measuring voltage, current, resistance, and possibly other electrical quantities and displaying the value in a digital format. Diode: A semi-conductor device that allows more current to flow in one direction than the other: ideally, it allows current to flow in one direction unimpeded, but allows no current to flow in the other direction. Dipole: An entity consisting of two equal and opposite poles is a dipole. There are electric dipoles and magnetic dipoles. An electric dipole consists of a positive charge and a negative charge separated by a distance (the distance is usually small compared to other distances being considered). A magnetic dipole is a magnetic system in which the equal (and opposite in character) north and 18-26

Glossary

south poles of a magnetic source are separated by a short but definite distance. A magnetic dipole will tend to orient itself parallel to an applied magnetic field in the same way an electric dipole does in an electric field. When dealing with two dimensions, an electric dipole consists of two infinitely long, parallel lines of charge (equal but opposite in sign), and a similar magnetic dipole consists of two infinitely long, parallel equal currents in opposite directions. A closed loop of current (either in a plane or along some irregular contour through space) comprises a three-dimensional dipole. Dipoles are field sources that attenuate with distance: the field decreases as a function of 1/r2 for two-dimensional dipoles (two parallel wires) and 1/r3 for three-dimensional dipoles (current loop). (See also Dipole Moment). Dipole Moment: A vector used to quantify the size of a dipole. The dipole moment is defined such that its vector cross product with an applied magnetic field gives the torque that the dipole experiences in the field. For a three-dimensional magnetic dipole (current loop), its dipole moment magnitude is the product of the current and the area enclosed by the loop, and in a direction given by a right hand rule. For a two-dimensional magnetic dipole (long parallel wires with equal and opposite current), the dipole moment is the product of its current and the distance between its two wires. (See also Dipole). Disc Signal: (Historical) A signal in which a colored disc is displayed behind a glass front in a closed case. Sometimes known as a banjo signal. Dispatching: A North American term referring to the act of putting a train into service from a station or depot or when they are reversed at a terminus. It was common for a dispatcher to allot crews to trains and issue dispatch orders to send the train out. The function is similar to the UK controller. Distant Field: The electric or magnetic field in the space located at distances such that the source of the electric or magnetic field can be represented within the desired accuracy by a simple element (e.g. a dipole or a quadrupole). For instance, the distant magnetic field, B, of a three-phase line with flat configuration, phase spacing, P, and balanced current, I, is expressed by the dipole equation B = 2 3 PI R 2 with an accuracy greater than 10% for distances R > 3.5 P. (B in milligauss, I in Ampere, P and R in meter.) Distant Source: A distant source is located far enough away from a shielding enclosure (or region of interest) for its field to be essentially uniform over the spatial region of interest. DMM: See Digital Multimeter. Domain: A microscopic magnetic region within a specimen composed of a group of atoms whose magnetic fields are aligned in a common direction. The size of a domain depends on various conditions, but usually contains millions of atoms. In some substances the shape appears to be like a long, slender rod with a transverse dimension of microscopic size, but with a length on the order of a millimeter or so. Domain Capture Effect: The presence of a large dc field tends to reduce the permeability of a shield material for a smaller ac field. Shielding problems involving ducting or flux diversion must consider both the external 60 Hz field from power lines and other sources and the earths static field. 18-27

Glossary

If the static geomagnetic field is larger than the 60 Hz field, the magnetic domains in the shielding material may be affected or captured by the flux from the earths magnetic field. Where the flux density from the earths field in the shield is larger than the ac flux density in the same position in the shield, the materials permeability for the 60 Hz field can be significantly reduced by the domain capture effect. Even if the earths magnetic field is not present or significant, the residual induction or retentivity (from some previous magnetization) may capture the domains and reduce the useful permeability. (See also Domain and Ducting). Double Break: A term used to describe the method of design in which both the positive and negative wires to the controlled device is opened by contacts of controlling device(s). Double Circuit Power Lines: Two separate overhead three-phase power lines using the same conductor support structures along some of their length. Dragging Detector: A device designed to alert a railway control center that a train has become derailed or a part of the train is dragging on the right of way. Normally consists of a bar placed across the four-foot (between the rails) that will be broken if struck. Often used on remote heavy haul freight lines where it is difficult for the crew to detect a mishap of this sort. Dual Control: (Railroad) A term applied to signal appliances provided with two authorized methods of operation. (See also Switch, Semi-Automatic Control). Dual Voltage Locomotive: (UK) A locomotive or multiple unit train designed to operate over lines having two different electric traction power supply systems. Locomotives have been designed to operate with up to four different voltages covering both ac and dc systems. Some trains can operate on lines with either overhead or third rail current collection as for the Eurostar Trains and UK Class 92 Channel Tunnel locomotives. Duct: A structurally solid enclosure with channels to provide a path for underground cables or conductors. Ducting: The shunting or redirection of magnetic flux lines to a low reluctance (high permeability) path in a magnetic circuit is called Ducting or Flux Shunting. A region can be shielded from magnetic fields by providing an alternate path around it for a portion of the flux. (See also Domain Capture Effect, Permeability, and Reluctance). Dynamic Braking: A method of braking in which the motor is used as a generator and the kinetic energy of the apparatus is employed as the actuating means of exciting a retarding force. (See also Brake Application). Earth Grounding System: A network of electrically interconnected rods, plates, mats, or grids installed for the purpose of establishing a low resistance contact with remote earth. Eddy Currents: When conducting specimens are subjected to a time-varying magnetic field (or motional induction in a static field); currents tend to be induced in the specimen. They flow in closed paths (e.g. circular paths perpendicular to the inducing field) in the specimen and are called eddy currents. In accordance with Lenzs law, the eddy current tends to oppose the change in the field inducing it. These induced currents will have a magnetic field associated with them. Eddy 18-28

Glossary

currents result in joule heating (I2R) in the conducting specimen. In fact, the technique of induction heating is an application of heat resulting from induced eddy currents. (See also Lenzs Law and Reflection). Effective Joint: (Railroad) The furthest joint (in the direction of travel) of a staggered pair of insulated rail joints, beyond which a shunt across the rails can become effective at the entering end of a track circuit. EHV: Extra High Voltage. Transmission system voltage between 300 kV and 800 kV. In the U.S. this includes 345 kV (nominal, 363 kV maximum system voltage), 500 kV (nominal, 550 kV maximum system voltage), 765 kV (nominal, 800 kV maximum system voltage). EIA: Electronic Industries Association (Formerly RMA). Electric Charge: The quantity of electricity that can be stored on a body. It is measured in coulomb (C), which is the charge flowing in one second in the cross-section of a circuit with the current of one ampere. Charges may be either positive or negative. The smallest charge is that of an electron and is equal to about 1.6 10-19 coulomb. Electric Field: An electric field exists in the region near electric charges and the field exerts a force on other electric charges placed in the field. At a given point in space, the ratio of force on a positive test charge (placed at the point) to the magnitude of the test charge, in the limit that the magnitude of the test charge goes to zero is defined as the electric field. The electric field strength (E-field) at a point in space is a vector defined by its space components along three orthogonal axes. For steadystate sinusoidal fields, each space component is a complex number or phasor. The magnitudes of the components are expressed by their rms values in volts per meter (v/m) or kilovolts per meter (kV/m) (emf per unit length). In a multi-phase environment, such as near a three-phase electric power line, the field is characterized as a vector rotating in a plane where it describes an ellipse whose semimajor axis represents the magnitude and direction of the maximum value of the electric field, and whose semi-minor axis represents the magnitude and direction of the field a quarter cycle later at its minimum value. Electric Field-Vertical Component: The vertical component of the electric field under a transmission line is the rms value of the component of the electric field along the vertical line passing through the point of measurement. This quantity is often used to characterize induction effects in objects close to ground level. (See also Electric Field). Electric Flux: The electric flux through a surface is the integral of the normal component of the electric flux density over the surface. Electric Flux Density: The electric flux density, D, is a vector quantity determined by the direction and density of flux lines that emanate from electric charge and follow the direction of the electric field. It is measured in coulomb per square meter. The electric flux density is related to the electric field by: D = E, where is the dielectric constant, or permittivity. Electric Locking: (Railroad) The combination of one or more electric locks and controlling circuits by means of which levers of an interlocking machine are locked, or the equivalent using circuits only, so that switches, signals, or other units operated in connection with signaling and interlocking, are secured against operation under certain conditions.

18-29

Glossary

Approach: Electric locking effective while a train is approaching, within a specified distance, a signal displaying an aspect to proceed, and which prevents, until after the expiration of a predetermined time interval after such signal has been caused to display its most restrictive aspect, the movement of any interlocked or electrically locked switch, movable point frog, or derail in the route governed by the signal, and which prevents an aspect to proceed from being displayed for any conflicting route. Detector Locking: A method of locking which is effective when the detector track circuit is occupied. Detector Locking prevents the operation of any power-operated switch, movable point frog or derail and the display of any signal indication more favorable than Proceed at Restricted Speed within the limits of the detector track circuit. Detector Locking is also referred to as Section Locking. Indication: Electric locking which prevents manipulation of levers that would result in an unsafe condition for a train movement if a signal, switch, or other operative unit fails to make a movement corresponding to that of its controlling lever, or which directly prevents the operation of a signal, switch, or other operative unit, in case another unit which should operate first fails to make the required movement. Route: Electric locking, effective when a train passes a signal displaying an aspect for it to proceed, which prevents the movement of any switch, movable point frog, or derail in advance of the train within the route entered. It may be so arranged that as a train clears a track section of the route, the locking affecting that section is released. (See also Sectional Release). Section: Electric locking effective while a train occupies a given section of a route and adapted to prevent manipulation of levers that would endanger the train while it is within that section. Time: A method of locking, either mechanical or electrical, which, after a signal has been caused to display an aspect to proceed, prevents, until after the expiration of a predetermined time interval after such signal has been caused to display its most restrictive aspect, the operation of any interlocked or electrically locked switch, movable point frog, or derail in the route governed by that signal, and which prevents an aspect to proceed from being displayed for any conflicting route. Traffic: Electric Locking that prevents the manipulation of levers or other devices for changing the direction of traffic into a section of track on which a route is lined, occupied, or locked. Electric Switch Lock: (See Lock). Electrical Conduit: Metal pipes containing two or more electrical wires. The conduits are often made of aluminum and are of circular section. They are at ground potential and in normal operation do not carry currents. In normal conditions the net current of an electrical conduit should be zero and the conduit will not produce any significant magnetic field. However, for a variety of reasons usually corresponding to accidental short circuits or miswiring that are violations of the National Electric Code, an electrical conduit may have a net current and this may be a significant source of power frequency magnetic field. 18-30

Glossary

Electrical Panel: A metallic box containing circuit breakers and electrical wires used for the distribution of electric power in a building. A building may have a Main Distribution Panel (MDP) where the electrical service entrance is located, and several sub-panels that receive electrical power from the MDP. Electro-Diesel Locomotive (UK): A locomotive that can operate with either electric or diesel power. The motors are electric but can be supplied from the diesel driven alternator or from the lines external power supply. The preference is usually for collecting electric power. The South African Class 38 is a locomotive of this type that has 3000v dc overhead electric supply as well as a 780 kW diesel engine. The UK Class 73 is another type but it collects electric power at 750v dc from a 3rd Rail to provide about 1500 h.p. or 600 h.p. with the diesel engine. The Swiss RhB railway also had two electro-diesel locomotives. Electromagnet: An electric current in a wire conductor has an associated magnetic field, but if the wire is wrapped in the form of a coil, the current and its magnetic field become concentrated in a smaller space inside the coil, resulting in a stronger field. The coil acts like a bar magnet (or permanent magnet), with opposite poles at the ends. More current and more turns make a stronger magnetic field. Sometimes, a ferromagnetic core is used to concentrate magnetic flux lines inside the coil. Soft iron is generally used for the core because it is easily magnetized and demagnetized. This ability of an electromagnet to provide a strong magnetic force of attraction that can be turned on or off easily has many applications in lifting magnets, buzzers, bells or chimes, and relays. A relay is a switch with contacts that are opened or closed by an electromagnet. Another common application is magnetic tape recording. The tape is coated with fine particles of iron oxide. The recording head is a coil that produces a magnetic field in proportion to the current. As the tape passes through the air gap of the head, small areas of the coating become magnetized by induction. On playback, the moving magnetic tape produces variations in electrical current. Electromagnetic Compatibility (EMC): The ability of a device, equipment, or system to function satisfactorily in its electromagnetic environment without introducing intolerable electromagnetic disturbances to anything in that environment. Electromagnetic Field: A time-varying field, with electric and magnetic field components described by Maxwells equations. Electromagnetic Interference (EMI): The impairment of the performance of equipment or systems caused by an unwanted electromagnetic disturbance. Electromagnetic Pulse (EMP): A transient high-intensity electromagnetic field. EMP is commonly associated with nuclear explosions in or near the Earths atmosphere; however, electromagnetic pulses can result from other sources, such as lightning. Electromagnetic Susceptibility: The level of electromagnetic interference that will cause a device, equipment, or system to malfunction or be damaged due to its electromagnetic environment. Electromotive Force: The term electromotive force (emf) is used to describe a potential difference (called voltage), indicating the ability to do the work of forcing electrons to move. The term emf as an abbreviation of electromotive force is sometimes used for a voltage source or an induced voltage. The unit of emf is the joule/coulomb, which is the volt. 18-31

Glossary

Electropneumatic Relay: A relay, the contacts of which are operated by air pressure. Electropneumatic Switch: (Railroad) A track switch operated by an electro-pneumatic switchand-lock movement. Electropneumatic Valve: A valve electrically operated which, when operated, will permit or prevent passage of air. Electrostatic: (See Quasistatic). ELF: Extremely-low-frequency. The frequency range from 3 to 3,000 Hz (per IEEE). EMC: (See Electromagnetic Compatibility). emf: (See Electromotive Force). EMF: A non-scientific term in popular use to refer to Electric and Magnetic Fields (or, Electromagnetic Fields, and sometimes used to refer only to the magnetic field); not to be confused with the engineering term for electromotive force (EMF). EMF Modeler: EMF Modeler is an EPRI computer program that combines the features and function of two earlier EPRI software packages: SUBCALC and RESICALC. This new combined program is a full 32-bit application designed for the Microsoft Windows95 or NT operating environments. This exceptionally versatile software includes new features based on user suggestions after significant work with SUBCALC and RESICALC. The most current version of the EMF Modeler is 1.0 and it is the central module in the EPRI EMF Workstation. (See also RESICALC, Power Line Calculator, and SUBCALC. EMF Workstation: The EPRI EMF Workstation (EMFW) provides a software tool for engineers to model electric and magnetic fields (EMF) produced by electric power lines and substations as well as exposure to the fields. The EMFW combines several EPRI EMF programs, reference materials, and utilities/tools into a single workstation environment. EMFW consists of the following modules: EMF Modeler, RESICALC, SUBCALC, EXPOCALC, ENVIRO, EMF Expert and the Power Line Calculator. Each module in the EMF Workstation includes a complete on-line help system, and the applicable modules share a common database. The TM software is compatible with PCs running the Microsoft Windows or NT operating systems. EMI: (See Electromagnetic Interference). EMTP: Electromagnetic Transient Program. Energized: A conductor, current carrying component, or other object that is electrically connected to a source of potential difference or electrically charged so as to have a potential significantly different from that of nearby earth. Engine: (Railroad) A unit propelled by any form of energy or a combination of such units operated from a single control, used in train or yard service. EP: Electropneumatic.

18-32

Glossary

EPRI: Originally called the Electric Power Research Institute, EPRI is a nonprofit organization committed to advancing electric power generation, delivery, and use throughout the world. Created by U.S electric utilities in 1973, EPRI is one of Americas oldest and largest research consortia, with some 700 members and an annual budget of about $500 million. EPRI is organized into 5 technical groups: Customer Systems (CSG), Environment (EG), Generation (GG), Nuclear Power (NPG), and Power Delivery (PDG). EPRI also maintains core competency in Strategic Research and Development (SR&D). EPRI today maintains a presence in North America, Europe, Asia, South Africa, and Latin America. Membership is open to any organization worldwide that generates, transmits, or distributes electricity. EPRI offers more than 80 scientific research and technology development areas. Equalizer (Lightning Protection): A device for equalizing voltage between two or more wires in the event of abnormal voltage rise between them. Equalizer (Line-To-Line Surge Protective Device): (Railroad) A two terminal primary SPD(s) for railroad signal applications which incorporates an ohmic leakage path to minimize static voltage buildup. This device provides metallic or normal mode protection. Equipment Ground: A ground connection to non-current-carrying metal parts of a wiring installation or of electric equipment, or both. Equipment Grounding Conductor: (Power) The conductor used to connect the non-current carrying metal parts of equipment, raceways, and other enclosures to the service equipment, the service power source(s) ground, or both. Equipment Grounding Conductor: (Railroad) Conductors that connect the ground of electrical panels to metallic parts of electrical equipment that must be kept at ground potential. These conductors may consist of wires (ground wires, usually green or bare) or of conduits. Equipment grounding conductors are not intended to carry current, except when there is an accidental contact between an energized wire and a grounded conductive object. Equipotential Plane: A mass (or masses) of conducting material that, when bonded together, provide a low impedance to current flow over a large range of frequencies. Equipotential Workzone (AREA, SITE): A work zone (area, site) where all equipment is interconnected by jumpers, grounds, ground rods, and/or grids that will provide acceptable potential differences between all parts of the zone under worst-case conditions of energization. Escapement Crank: (See Crank) (Historical). Exceedance Level: The level of a quantity that is exceeded a specified percentage of times. For instance the 90% exceedance level or L90 is the level exceeded 90% of the time. Sometimes, however the opposite meaning, i.e. the meaning of non-exceedance level is given to the same notation. Therefore, L90 sometimes means a level that is exceeded only 10 percent of the time. In this book the meaning of exceedance level is maintained as first stated.

18-33

Glossary

Exothermic Welding: The process of joining metals by the reduction of copper oxide using aluminum. This chemical reaction produces heat along with a molten copper. The molten copper flows over the conductors being welded and provides a molecular homogeneous weld for grounding and bonding. Extremely-Low-Frequency: (See ELF). Extruded-Dielectric Cables: Extruded-dielectric cables use either crosslinked polyethylene (XLPE), low or high density polyethylene, or ethylene-propylene rubber for insulation. Extruded-dielectric cables are also called solid dielectric or XLPE cables. The reason for use of the extruded dielectric materials for insulation in high voltage cables is to provide a simple, low cost, low maintenance alternative to conventional pipe-type cable systems that use fluid or gas for insulation. These cables are fabricated as single-phase cables and usually placed in a duct with some spacing between the cables that results in less magnetic field cancellation than if the cables were all placed together. In addition, extruded-dielectric cable construction does not normally have a common shield around all phases, as is the case with the steel pipe in pipe-type cables. Facing Movement: The movement of a train over the points of a switch that face in a direction opposite to that in which the train is moving. Facing Point Switch: (Railroad) A track switch, the points of which face traffic approaching in the direction for which the track is signaled. Facing Points: These are points where a single track diverges into two, i.e. where the points face the oncoming train. The opposite is trailing points, where two tracks become one. Fail Safe: A term used to designate a railway signaling design principle, the objective of which is to eliminate the hazardous effects of a failure of a component or system. (See also Right-Side Failure and Vital Circuit). Failure: A term used when a device does not perform its intended function. False Activation: The activation of the highway-railroad grade crossing warning system in response to a condition which requires correction or repair of the warning system, either in the control or track circuit. This is a safe failure of the system, in that it provides notice of warning as is required by fail-safe design techniques. False Clear: (False Proceed) The failure of a signaling system. (See Wrong-Side Failure). False Proceed: (False Clear) A failure of a system, device or appliance to indicate or function as intended that results in less restriction than is required. (See Wrong-Side Failure). False Restrictive (FR): A failure of a system, device or appliance to indicate or function as intended that results in greater restriction than is required. False Restrictive Aspect: The aspect of a signal that conveys an indication more restrictive than intended. Far Field: The far field is a term used to denote the region far from a radio frequency (RF) antenna compared to the wavelength corresponding to the frequency of operation. In the far field region an electromagnetic wave generally has the characteristics of a plane-wave: it 18-34

Glossary

propagates at the speed of light in a direction perpendicular to its electric and magnetic fields. For a propagating (radiating) electromagnetic field, the electric field depends on the magnetic field through the time variation of the magnetic field, and in turn, the magnetic field depends on the electric field through the time variation of the electric field. Sometimes, the distance from an antenna beyond which the transmitted power density decreases inversely with the square of the distance. Farad: A unit electrical capacitance named after the British physicist Michael Faraday (1791-1867). A Farad is equal to the capacitance that carries a charge of one coulomb when it is charged by a potential difference of one volt. It is a rather large unit and, in practice, much -6 -12 smaller units such as microfarads (10 farad) or picofarads (10 farad) are generally used. (See also Capacitance). Faraday Shield: Electrostatic and electromagnetic shield composed of a metallic enclosure, wire mesh, or a series of parallel wires usually connected at one end to another conductor that is grounded. Also known as Faraday Cage or Faraday Screen. Faradays Law: The quantitative relationship between the emf (electromotive force) induced in a closed loop and the magnetic field producing the emf is given by Faradays Law. According to this law, the total emf induced in a closed circuit is equal to the time rate (dt) of decrease of the total magnetic flux () linking the circuit. emf: (volts) = -d/dt Fault: (Electrical) A physical condition that causes a device, a component, or an element to fail to perform in a required manner, for example, a short-circuit, a broken wire, an intermittent connection. Feed: (in electrical circuits) In an electrical circuit the feed is the electrical connection between the power supply and parts of the circuit. Feeder Circuit: (Power) In a power distribution system a feeder circuit is a three-phase power line connecting a distribution substation with several loads. The loads may be derived directly from the feeder or from other distribution lines connected to the feeders. For example, a feeder circuit may be located along a major road and several other lines may be derived from it to serve secondary streets. These lines, which may be three-phase, two-phase, or single-phase, are called laterals. Federal Railroad Administration (FRA): The FRA was created in 1966 to promote and enforce safety throughout the United States railroad system, rehabilitate the Northeast Corridor rail passenger services, consolidate Federal support for rail transportation, and support research and development for rail transportation. The FRA administers and enforces the Federal laws and related regulations designed to promote safety on railroads; exercises jurisdiction over all areas of rail safety under title 49, United States Code, chapter 201, such as track maintenance, inspection standards, equipment standards, and operating practices. Railroad and related industry equipment, facilities, and records are inspected and required reports reviewed. In addition, FRA educates the public about safety at highway rail grade crossings and the danger of trespassing on rail property.

18-35

Glossary

Ferrites: A group of ceramic materials having the ferromagnetic properties of iron (high permeability), but with high electrical resistance are called ferrites. Ferrites are iron oxides combined with oxides of other metals such as manganese, cobalt, nickel, copper, and magnesium. The combined oxides are powdered, formed into the desired shape under pressure, and fired. A 2 ferrite core is much more efficient in a coil when the current alternates very rapidly, since the I R power lost by eddy currents in the magnetic core is then less because of the high resistance. However, the ferrites are easily saturated at relatively low values of magnetizing current. Ferromagnetism: The property of certain materials with high values of permeability that become strongly magnetized in the same direction of an applied magnetizing field. In the atoms of ferromagnetic materials, there are unpaired electrons whose spins are oriented in the same way. The common metals iron, cobalt, and nickel; the rare earth elements gadolinium and dysprosium; some alloys of these and other elements; and certain metallic oxides called ferrites (i.e. iron oxides combined with oxides of other metals such as manganese, cobalt, nickel, copper, magnesium) show strong ferromagnetic properties. In these materials a special effect occurs which permits a specimen to achieve a high degree of magnetic alignment in spite of the randomizing tendency of the thermal motions of the atoms. An interaction (within a material) called exchange coupling occurs between adjacent atoms, coupling their magnetic moments together in rigid parallelism. If the temperature of a ferromagnetic material is raised above a certain critical value, called the Curie temperature, the exchange coupling suddenly disappears and the material simply becomes paramagnetic (for iron the Curie temperature is 770C). (See also Curie Temperature and Permeability). Field Blocking: (Railroad) Blocking which makes use of a vital relay located at the controlled point or remote controlled interlocking. (See also Office Blocking). Field Ellipse: An electric or magnetic field can have three orthogonal components in space that can each vary in time. The instantaneous magnitude and direction of the field at a point in space is defined by these components with proper consideration of the spatial and temporal aspects of each component. The field at a point can be represented in most cases by a field ellipse located in three dimensions. For many practical cases near power lines the field ellipse throughout a large region can be represented in two dimensions if the component of field parallel to the power line is small. The field ellipse can be characterized by its major and minor axis, the maximum and minimum values obtained by the ellipse (the field is said to be elliptically polarized). The maximum value of the field at a point is simply the value of the semi-major axis (one-half of the major axis dimension) and the minimum value is the semi-minor axis. The resultant is the square root of the sum of the squares of the semi-major and semi-minor axes. There are two special cases for the field ellipse. In one case the major and minor axes are equal and the field ellipse becomes a circle; the field is said to be circularly polarized. In the other case the minor axis goes to zero and the field is then represented by a space vector that has a magnitude that varies with time along a straight line; the field is said to be linearly polarized. (See also Polarization). Field Line: A line in an electric or magnetic field that shows the direction of force exerted by the field. (See also Flux). Field Management (Power) The application of engineering principles, based on research, to electric and magnetic fields due to power lines or other sources to address or treat situations requiring technical solutions without altering the function for which the system was intended.

18-36

Glossary

Filter: A circuit that selects or rejects one or more components of a signal related to frequency. Low Pass Code: A low pass filter connected between railroad coding equipment and line to pass direct current code impulses and prevent code equipment from shunting carrier and communication circuits. It prevents line-coding contacts from introducing undesired high frequency currents into the line. Finite Difference Method: A numerical technique to calculate electric or magnetic fields. The field values are derived from Maxwell differential equation. The method of finite differences applied to the solution of electric field problems solves for the value of the potential at discrete points of a regular mesh placed over the field region. The method is called finite differences because the partial derivative of the potential is replaced by the finite difference between the potentials at adjacent points of the mesh. Finite Element Method: A numerical technique to calculate electric or magnetic fields based on Maxwell differential equations. The finite element method applied to the solution of electric field problems is based on the principle that the equilibrium distribution of charges present in a closed domain is such as to minimize the electrostatic energy associated with the field. The whole domain is divided into a finite number of sufficiently small elements of simple geometry connected to each other to form a mesh. The electrostatic energy is written as a function of the potentials of the nodes of each element. The minimum energy is obtained by setting the partial derivatives of the total energy with respect to the potential of each node to zero. A system of as many equations as there are node potentials is obtained. Fixed Signal: (See Signal). Flasher: (See Relay). Flashing Light Signal: (See Signal). Flat Configuration: (Power) A configuration of a single-circuit three-phase line in which the three phases are all at the same height above ground (also called a horizontal configuration line). Fleeting: (Railroad) A facility which provides for automatic clearing of controlled signals in one direction for successive trains. Floating Slab Track (UK): A track system using a concrete base mounted on rubber pads to reduce vibration transmitted to adjacent property. A number of railways have tried it with some anecdotal evidence suggesting that the maintenance costs are high and that riding quality of trains suffers. Floor Push: (Railroad) (Historical) A circuit controller mounted in the floor so that a circuit may be made by pressing on a plunger. Fluorescent Light Ballast: (Power) (See Ballast-Lighting).

18-37

Glossary

Flux: A property of any vector field - it is represented by lines of force that cut through a reference surface. A line of flux is a line so drawn that a tangent to it at any point indicates the direction of the field. There are electric and magnetic flux lines. The electric flux lines are also called lines of force and the magnetic flux lines are sometimes called lines of induction. If a test charge were placed in an electric field, or an isolated pole in a magnetic field, the flux lines would describe the direction of force on the test element and hence, its direction of movement within the field. The unit of magnetic flux in the MKS (SI) system is the weber and the unit in the CGS system is the maxwell. The number of flux lines per unit cross-sectional area is proportional to the magnitude of the flux density vector (D for electric flux density and B for magnetic flux density). The flux lines give a graphic representation of the way a field varies throughout a certain region of space. Where the flux lines are close together, the field is large and where they are far apart, the field is small. (See also Magnetic Flux Density, Maxwell, and Weber). Flux-Entrapment Shields: A shield of high permeability material that provides a low reluctance path for magnetic field lines so that the flux is diverted (shunted or redirected) into the material away from the region to be shielded. (See also Ducting and Reluctance). Flux-Gate Magnetometer: The flux-gate magnetometer is an instrument that uses the flux-gate magnetic sensor technique for measurement of magnetic fields. The flux-gate probe consists of a ferromagnetic core wound with two coils (one for excitation and one for detection). The flux-gate technique utilizes magnetic induction and the hysteresis exhibited by all ferromagnetic materials. A high purity sinusoidal current is applied to the excitation coil to generate an alternating magnetic field in the core. This excitation current magnetizes the core, continuously driving it in and out of saturation. Because of hysteresis, the magnetic flux through the core will trace a loop if it is plotted against the magnetic field intensity. The detection coil senses changes in the flux density through the core. As the core is driven into saturation, the reluctance of the core to the external axial magnetic field being measured increases, making it less attractive for the external magnetic field to pass through the core. As the field is repelled, the detection coil senses its change in flux density. When the core comes out of saturation by reducing the current in the excitation coil, the external magnetic field is again attracted to the core, which is again sensed by the detection coil. Thus, alternate attraction and repulsion (gating in and out) causes the magnetic lines of flux to cut the detection coil and the induced current is converted to a linear voltage that is proportional to (and used to measure) the external magnetic field. Flux Shunting: (See Ducting). Follow Current: (Follow on Current) The current from the connected power source that flows through a surge protective device during and following the passage of discharge current. Foreign Current: (Railroad) A general term applied to stray electric current which may affect the signaling system, but which is not a part of the system. Forestall: (Train Stop, Train Control, etc.) A term meaning the prevention of an automatic brake application by operation of the acknowledging device or by manual control of the speed of train.

18-38

Glossary

Fouling Circuit: (Railroad) The track circuit in the fouling section of a turnout, connected in multiple with the track circuit in the main track. Fouling Point: The location on a turnout back of the frog at which insulated joints or derails are placed at or beyond clearance point. (See also Clearance Point). Fouling Section: The section of track between the switch points and the fouling point in a turnout. Four Foot (UK): Nickname for the gap between the running rails, the gauge. Comes from the standard gauge measurement of 4ft-8 1/2 inches (1435mm). Four Rail System (UK): A now almost unique current collection system used by the London Underground that has separate positive and negative current rails. The usual 3-rail method of connecting return current via the running rails is replaced by a fully insulated system using separate positive and negative rails. Originally used to reduce the risk of stray currents causing damage to nearby utilities and structures through electrolysis. It has the disadvantage of requiring special fault detection, as earth faults do not cause current to automatically switch off. See conductor rail. Free-Body Electric Field Meter: A type of electric field meter where the detector is housed between the two electrodes of the sensor and the whole assembly is electrically floating in space by means of an insulating handle without any electrical connection to ground. FPL: Facing point lock. FRA: Federal Railroad Administration. Frequency-Selective Voltmeter: A special voltmeter with a narrow band input filter that allows measurement of a single, user-selectable frequency component of a complex signal. Frog: A railroad track structure used at the intersection of two running rails to provide support for wheels and passageways for their flanges, thus permitting wheels on either rail to cross the other. Fused Disconnect: A device used to connect distribution transformers to distribution line phase wires. Operable with a hot stick, they provide the power company a means to connect and disconnect the load. Under fault conditions the fuse element within the device melts and the device opens. Galloping: The large amplitude, low-frequency, wind-induced vibration of electric power conductors, usually caused by icing and steady, moderate wind conditions (~5-45 mph). Gamma: A smaller unit of magnetic field strength in the CGS electromagnetic system is the -5 gamma (). One gamma equals 10 oersted. The earths magnetic field has sometimes been stated in units of gamma. (See also Oersted).

18-39

Glossary

Gap: (Railroad) A break in the continuity of a conductor rail at a set of points on a line equipped with 3rd rail traction. Gas Insulated Substation (GIS): For applications where space requirements are a problem; conventional bus arrangements are replaced with gas insulated substations. The gas is generally SF6. Each conductor is placed in the center of an enclosure filled with SF6, which exhibits excellent dielectric strength and, therefore, allows small distances between the energized conductor and the enclosure, which is at ground potential. Gauge: (of Track) The distance between the gauge lines, measured at right angles thereto. (The standard gauge in North America is 4 ft. 8-l/2 in.) Gauss: The gauss is a unit (in the CGS system) used to describe magnetic flux density (B), or magnetic flux lines () per unit of cross-sectional area. The gauss is one maxwell (one flux line) per square centimeter, or 10-4 webers per square meter. One gauss equals 10-4 tesla (MKS or SI units) and 1 mG equals 0.1 T. (See also Magnetic Flux Density, Maxwell, Tesla and Weber). Gate: (Railroad) (See Highway Grade Crossing Gate). Geometric Mean Diameter (GMD): The geometric mean diameter of a conductor is used to characterize its diameter for inductance calculations. The GMD of various conductors is standard information usually provided by the conductor manufacturer. The EPRI Transmission Line Reference Book/345 kV and Above provides tables of conductor data that include the GMD and also formulas to estimate the GMD for single or bundled configurations. The GMD of a conductor(s) is used in calculations related to active and passive shielding. GFCI: Ground Fault Circuit Interrupter. A device designed to detect the net current of a pair of wires (e.g. hot and neutral) and trip the breaker. GFI: Ground Fault Interrupter. A device designed to prevent shocks by opening a circuit at a location other than the electrical panel, near the point where a ground fault may occur. GIS: Gas Insulated Substation. Grade: A North American term for sloping track (UK terms are gradient or bank). The phrase at grade means level track. Gradient: (Railroad) (See Grade). Ground: A conducting connection, whether intentional or accidental, by which an electrical circuit or equipment is connected to the earth, or to some conducting body of relatively large extent that serves in place of the earth. Ground Bus: An electrical bus to which the grounds from individual pieces of equipment are connected, and that, in turn, is connected to ground at one or more points. Ground Current: Current flowing in the earth or in a grounding connection.

18-40

Glossary

Ground Rods: (Power) Copper rods driven into the earth for about 8 feet. They are used as grounding electrodes for residential services and are intended to provide a ground impedance of 25 or lower. Grounded: A conductor, current carrying component, or other object that is intentionally connected to earth through a ground connection or connections of sufficiently low impedance and having sufficient current-carrying capacity to limit the build up of voltages to levels below that which may result in undue hazard to persons or to connected equipment. Grounding Bus (or Bar): The grounding bus of an electrical panel is the electrode to which all the equipment grounding conductors of the electrical circuits of the panel are connected. The grounding bus is connected to the metallic case of the panel. Grounded Circuit: A circuit in which one conductor or point (usually the neutral conductor or neutral point of transformer or generator windings) is intentionally grounded, either solidly or through a grounding device. Also, a circuit in which one conductor or point is accidentally grounded, either directly or through a grounding device. Grounded Conductor: A system or circuit conductor that is intentionally grounded. Ground-Fault Circuit-Interrupter: A device intended for the protection of personnel that functions to interrupt the electric current to the load within an established period of time when a fault current to ground exceeds some predetermined value that is less than that required to operate the over current protective device of the supply circuit. Ground Grid (Ground Electrode System): A system of grounding electrodes consisting of interconnected bare cables buried in the earth to provide a common ground for electrical devices and metallic structures. Ground Potential Rise (GPR): The difference in ground potential between a location in proximity to a point of large current injection into the ground and any remote ground point. GPR is usually caused by a short circuit of an energized power conductor to ground and is the result of the injected current flowing through the impedance of the ground circuit. Lightning can also cause GPR. Grounding Conductor: A conductor used to connect equipment or the grounded circuit of a wiring system to the grounding electrode or electrodes. Grounding Electrode: (Power) An electrode in electrical contact with the earth. An approved grounding electrode must have a resistance to ground of 25 or lower. Commonly used grounding electrodes are ground rods, the steel of steel-reinforced concrete buildings, and metallic water pipes, provided that these pipes are in the earth for 10 feet or more. Grounding Electrode (Earth Electrode): A conductor used to establish a ground, through direct contact with the soil. Grounding Wire: (Power) The conductor connecting the neutral at the service entrance to the grounding electrode. It is usually a heavy conductor connected to a clamp on a ground rod or a water pipe.

18-41

Glossary

Guy Wire: A high strength cable used to provide mechanical support for power line structures. Hand Operated Switch: (Railroad) A non-interlocked switch that can only be operated manually. Harmonic Content: Harmonic content is the distortion of a sinusoidal waveform characterized by indication of the magnitude and order of the Fourier series terms describing the wave. For power-frequency fields, the harmonic content of the electric field coincides with that of the line voltage, and the harmonic content of the magnetic field coincides with that of the line current for single-phase systems. For power transmission lines, the harmonic content is small, except during transient conditions, and of little concern for the purpose of field measurements except at points near large industrial loads such as saturated power transformers, n-pulse rectifiers, or aluminum, chlorine, and graphite plants where certain harmonics may reach 10% of the line voltage. Laboratory installations may also have voltage or current sources with significant harmonic content. (See also Total Harmonic Distortion). Harmonics: Integer multiples of a fundamental frequency are called harmonics (or overtones). The lowest frequency is called the fundamental frequency; for example the fundamental frequency of the North American electric power system is 60 Hz (other countries such as the UK have a fundamental frequency of 50 Hz). For a 60 Hz fundamental, the second harmonic would be 120 Hz, the third harmonic would be 180 Hz, and so forth. Head Block Signal: (Railroad) A home signal governing entrance into the block between sidings on single track. Helmholtz Coils: A term used to describe a coil system that can produce nearly uniform magnetic fields over a significant volume for calibration purposes. Helmholtz coils have many turns of wire and are identical in construction; the coils can have either circular or rectangular geometry. The pair of coils are placed parallel to each other, spaced at one-half their diameter apart, and are energized by a common source. Careful construction of the coils to ensure accurate dimensions is critical to producing uniform fields. An accurate method of determining the current supplied to the coils is necessary. The measurement of coil current usually employs a precision component or method that is traceable to NIST to guarantee accuracy. H-Frame: (Power) An electric power structure generally constructed with two vertical supports that are connected by a horizontal member, thus forming the shape of an H. High-Phase Order Transmission Line: Transmission lines with more than three phases per circuit. The phases are multiple of three. Six-phase and Twelve-phase transmission projects have been the object of several studies and of at least one electric company pilot project. They utilize the space more efficiently than three-phase transmission. The space compaction results in reduced electric and magnetic fields. Highway Grade Crossing: An intersection of a highway with a railroad track at the same elevation.

18-42

Glossary

Highway Grade Crossing Gate: A device that forms part of a Highway Grade Crossing Warning System that provides a visual warning and restricts access to the intersection of a Highway Grade Crossing. Highway Grade Crossing Sign (Crossbucks): A sign located at the intersection of a railroad-highway crossing at grade to warn highway traffic of the intersection. Highway Grade Crossing Signal: That part of a Highway Grade Crossing Warning System used at the crossing that provides the visual warning to highway traffic. (See also Signal- Application, Flashing-Light). Highway Grade Crossing Warning Device: (See Bell, Gate, Signal). Highway Grade Crossing Warning System: An interconnection of various devices and their controls used to indicate the approach and/or presence of a train at a highway grade crossing. Hold Clear: A term used to designate a device for holding a signal in any position other than its most restrictive. Home Signal: (See Signal). Horn: (Railroad) Electrically or pneumatically operated warning device provided for a driver to sound at will. Replaces the traditional whistle. Hot Box: A term for an axle box that has become over heated because of a breakdown of lubrication or excessive overloading. Hot Box Detector: (See Detector). Hot Leg (of a Residential Service): One of either of the pair of energized conductors servicing a residence. Two energized phase conductors and a neutral conductor typically provide service from a transformer (in other countries this can be three phase service). Each energized conductor is called a hot leg and has a nominal voltage of 120V with respect to ground or 240V with respect to the other hot leg (the pair of hot legs are 180 electrical degrees apart). Housing: An equipment enclosure (also called a bungalow; in the UK called a shed). HPFF: High Pressure Fluid Filled. (See Pipe-type cable). HPGF: High Pressure Gas Filled. (See Pipe-type cable). HPOF: High Pressure Oil Filled. (See Pipe-type cable). HST: High-speed train. Hump Yard: A railroad classification yard in which the sorting of cars is accomplished by pushing them over a summit, known as a hump, beyond which they run by gravity and are switched into selected tracks.

18-43

Glossary

HVDC: High Voltage Direct Current. Hysteresis: The term hysteresis (a lagging behind) means the dependence of the state (or condition) of a physical system on its previous history. For a specimen placed in a magnetic field, it means that the specimens previous magnetization history will affect its present state. Hysteresis results from the fact that the materials magnetic dipoles are not perfectly elastic. Once aligned by an external magnetizing force, the dipoles do not return exactly to their original positions when the force is removed. If the magnetic field, H, applied to a specimen is increased to saturation and is then decreased, the magnetic flux density, B, decreases, but not as rapidly as it increased along the initial magnetization curve. If H is further increased in the negative direction, the specimen becomes magnetized with the opposite polarity, the magnetization at first being easy and then hard as saturation is approached. The phenomenon (or material property) which causes B (flux density) to lag behind H (field), so that the magnetization curve for increasing and decreasing fields is not the same, is called hysteresis, and the loop traced out by the magnetization curve (B-H diagram) is called a hysteresis loop. If the specimen is carried to saturation at both ends of the magnetization curve, the loop is called the saturation, or major hysteresis loop. The residual flux density Br on the saturation loop is called the retentivity, and the coercive force Hc on this loop is called the coercivity. Thus, the retentivity of a substance is the maximum value that the residual flux density can attain and the coercivity the maximum value that the coercive force can attain. For a given specimen no points can be reached on the B-H diagram outside the saturation hysteresis loop, but any point inside can. In soft (easily magnetized) materials the hysteresis loop is thin, with a small area enclosed. The hysteresis loop of a hard magnetic material is much wider with a greater enclosed area. (See also Magnetization Curve and Nonlinear Materials). Hysteresis Loss: When the magnetizing force applied to a material rapidly reverses direction, as with an alternating current, the material hysteresis can cause a considerable energy loss. The hysteresis results from the fact that the magnetic dipoles are not perfectly elastic. Once aligned by an external magnetizing force, the dipoles do not return exactly to their original positions when the force is removed. (The effect is the same as if the dipoles were forced to move against an internal friction between molecules.) A large part of the magnetizing force may then be used just for overcoming the internal friction of the molecular dipoles. The work done by the magnetizing force against this internal friction produces heat. This energy wasted in heat as the molecular dipoles lag the magnetizing force is called hysteresis loss. For steel and other hard (not easily magnetized) magnetic materials, the hysteresis losses are much higher than in soft (more easily magnetized) magnetic materials like iron. Frequency affects hysteresis loss because the faster the magnetizing force changes, the greater the hysteresis effect. (See also Hysteresis). ICC: Interstate Commerce Commission (United States). ICEA: Insulated Cable Engineers Association. IEEE: The Institute of Electrical and Electronics Engineers, or IEEE (formerly AIEE), is the worlds largest technical professional society. Founded in 1884, IEEE is comprised of more than 320,000 members who conduct and participate in its activities in approximately 150 countries. The technical objectives of the IEEE focus on advancing the theory and practice of electrical, electronics and computer engineering and computer science. IEEE promulgates standards that apply to many subjects, including EMF measurements, instrumentation, and calibration. 18-44

Glossary

Image Conductors: An electric field calculation technique consists of replacing the earth or ground plane (modeled as a plane at zero potential) with a set of conductors, called image conductors. The image conductors are located as the geometric mirror image of the phase conductors (but with opposite sign charges) below the ground plane. A somewhat similar use of image conductors can be used for magnetic fields but since the earth is not a perfect conductor the depth of the images below the ground plane is independent of the height of the phase conductors and, for a homogeneous earth, is approximately equal to 660 f (m). For example for = 100 m and f = 60 Hz. The image depth is equal to 852 m. Immunity: The ability of electronic equipment or systems to perform satisfactory in the presence of a specified electromagnetic environment. Impedance Bond: An iron core coil of low resistance and relatively high reactance, used on electrified railroad to provide a continuous path for the return propulsion current around insulated joints and to confine the alternating current signaling energy to its own track circuit. Impulse: A surge of unidirectional polarity. Impulse Sparkover (Flashover) Voltage: The highest magnitude of voltage attained by an impulse of a designated wave shape and polarity applied across the terminals of a surge arrester prior to the flow of discharge current. Sometimes referred to as surge or impulse breakdown voltage. Impulse Transformer: A transformer whose primary is connected in series with a direct current code line circuit and whose secondary actuates a device, usually a polar relay, in response to changes of the code line current. Impulse Withstand Voltage: The crest value of an applied impulse voltage that, under specified conditions, does not cause a flashover, puncture, or disruptive discharge on the test specimen. In Advance of a Signal: (Railroad) A term used in defining the territory beyond a signal as seen from an approaching train. In Approach of a Signal: (Railroad) A term used in defining the territory to which a signal indication is conveyed. In Rear of a Signal: (See In Approach of a Signal). Indication: (Signal) The information conveyed by the aspect of a signal. Indication Lock (Electric): (Railroad) (Historical) (See also Lever Indication) An electric lock connected to a lever of an interlocking machine to prevent the release of the lever or latch until the signals, switches or other units operated, or directly affected by such lever, are in the proper position. Indication Locking: (See Electrical Locking). 18-45

Glossary

Induction: The process of generating time-varying voltages and/or currents in conductive objects or electric circuits by the influence of the time-varying electric, magnetic, or electromagnetic fields. In-Span Spacers: Insulating spacers located within a span or spans of a power line and installed between phases to maintain adequate mechanical and electrical clearances are called in-span spacers. These spacers are often used when line phase spacing is to be compacted or to minimize damage due to wind-induced conductor motion. Instrument: (Railroad) (Historical) See Track Instrument. Insulated Rail Joint: A joint in which electrical insulation is provided between adjoining rails. (See also Effective Joint). Insulation Resistance: The resistance offered by the insulation on any current-carrying part or conductor. Interlocked Switch: (Railroad) A track switch which is interlocked so as to prevent the switch from being improperly moved or positioned in an unsafe manner. Interlocking: (Railroad) An arrangement of signals and signal appliances so interconnected that their movements must succeed each other in proper sequence and for which interlocking rules are in effect. It may be operated manually or automatically. Automatic: An arrangement of signals, with or without other signal appliances, which functions through the exercise of inherent powers as distinguished from those whose functions are controlled manually, and which are so interconnected by means of electric circuits that their movements must succeed each other in proper sequence, train movements over all routes being governed by signal indication. Manual: An arrangement of signals and signal appliances operated from an interlocking machine and so interconnected by means of mechanical and/or electric locking that their movements must succeed each other in proper sequence, train movements over all routes being governed by signal indication. Interlocking Limits: The tracks between the extreme opposing home signals of an interlocking. Interlocking Machine: (Railroad) (Historical) Electric: (Historical) An interlocking machine for the control of electrically operated functions. Electro-Mechanical: (Historical) An interlocking machine for the control of both power and mechanically operated functions. Electro-Pneumatic: (Historical) An interlocking machine for the control of electropneumatically operated functions. 18-46

Glossary

Mechanical: (Historical) An interlocking machine for the control of mechanically operated functions. Table: (Historical) An interlocking machine for the control of power-operated functions and designed for mounting on a table or desk. Interlocking Relay: (Railroad) (Historical) A relay having two independent magnetic circuits with their respective armatures so arranged that the dropping away of either armature prevents the other armature from dropping away to its full stroke. Interlocking Signals: (Railroad) The fixed signals of an interlocking. Interlocking Station: (See Station). Intermittent Control: (Cab Signal, Train Control, etc.) (See Continuous Control) A type of control in which the locomotive apparatus is affected only at certain designated points, usually at railroad signal locations. Intermittent: A temporary or unpredictable event. Inverter: (Railroad) Electronic power device mounted on trains to provide alternating current from direct current. IRSE: Institution of Railway Signal Engineers. Isolation: The physical and electrical arrangement of the parts of an equipment, system, or facility to prevent uncontrolled electrical contact within or between the parts. Isolation Transformer A transformer of the multiple-winding type, with the primary and secondary windings physically separated. This isolation separates grounded ac circuits from floating/ungrounded ac circuits. Joule: A unit measure of energy. The work done when the point of application of a force of one Newton is displaced a distance of one meter in the direction of the force (1 joule = 1 wattsecond). Jumpers: Multi-core cables used to provide electrical connections between railway vehicles. Knife Switch: (Railroad) A standard electric power isolating system mounted on the underframe of an electric vehicle to disconnect power and auxiliary systems from the traction current collection system and provide a link to shore (external) supplies. Kelvin Connection: The Kelvin or four-wire connection is commonly used in precision measurement systems to obtain accurate measurements of impedance. L5: The level exceeded 5% of the time. 18-47

Glossary

L10: The level exceeded 10% of the time. L50: The level exceeded 50% of the time. L90: The level exceeded 90% of the time. L95: The level exceeded 95% of the time. Lateral Lines: (Power) (See Feeder Circuit). Lateral Profile: The term lateral profile describes a series of measurements used to evaluate the spatial characteristics of a field as a function of distance away from a source. The most common example is a transmission line electric or magnetic field lateral profile, where the field is measured as a function of distance away from a transmission line, usually at equal spacing between measurement points. Typically, the lateral profile will be initiated next to the source (or directly underneath of a power line) and proceed away from the source in a direction perpendicular to the source, taking field readings at specific incremental distances. There is an IEEE Standard for measurement procedures for ac power lines (IEEE Std. 644-1994). Lead: North American term for a track giving access from a main line to a railway yard. Leakage Current: An electric current of relatively small value that flows through unintended paths or across the surface of insulation when a voltage is impressed across the insulation. Leakage Current- Ballast: (Railroad) The leakage current from one rail of a track circuit to the other through the ballast, ties, etc. Leave Siding Indicator: An indicator used to convey instruction for a train to leave the siding. (See also Take Siding Indicator). LED: Light Emitting Diode. Lenzs Law: The current in a conductor as a result of an induced voltage (emf) is such that the change in magnetic flux due to it is opposite to the change in flux that caused the induced voltage. Thus, Lenzs Law states that the induced current produces magnetic flux that opposes the change in the inducing field. (See also Eddy Currents). Let Go Current: An electric current flowing through a persons hand of sufficient magnitude that the person loses control of the hand muscles and cannot release the grip on an energized object. Lever Blocking Device: A device for blocking a lever so that it cannot be operated. Light Rail Vehicle (LRV): Modern generic term for tram or streetcar. An electrically powered rail vehicle using rails embedded in the roadway or using dedicated rail tracks, or a combination. LRVs are popular as an alternative to subway or underground rail lines for urban rail systems due to their reduced construction costs. 18-48

Glossary

Lightning Arrester: (See Surge Protective Device) A device for protecting circuits and apparatus against lightning or other abnormal potential rises of short duration. Air Gap: An arrester consisting of a spark gap between terminal plates in which the discharge is allowed to spread over a number of minute discharge points. Gas Tube: An arrester consisting of one or two gaps between electrodes maintained in a sealed tube containing a special gas or partial vacuum. Valve Type: An arrester containing a valve element that, because of its nonlinear currentvoltage characteristic, limits the voltage across the arrester terminals during the flow of discharge current and contributes to the limitation of follow-through current after the voltage surge has dissipated. Lightning Surge: A transient electrical disturbance in an electric circuit caused by lightning. Like Phasing: (See Super Bundle Phasing). Line Capacity (UK): (Railroad) The maximum number of trains capable of being operated over a line in one direction. Usually expressed as trains per hour, it will depend on all trains running at the same speed, having equal braking capacity and on how the signaling is arranged. Line Charge: (Power) An electrical charge distributed uniformly along a straight line. Used for calculations of the electric field on conductor surfaces, corona effects, and electric field in proximity of power lines. The unit of measurement is C/m (coulomb per meter). The following equations relate the line charge, q, to the average electric field on the surface of a cylindrical conductor, G, to the electric field, E, and to the space potential, Vsp, at a point in space (if no other charges were present): , where R is the distance between 2 R 2 R conductor center and measuring point; the field is directed from the conductor to the measuring point.

G=

, where R is the conductor radius. E =

q ln (R Ro ) , where R is the distance between conductor center and measuring point and 2E Ro is the distance between conductor center and a reference point where the space potential is set to zero. Vsp =
Line Circuit: (Railroad) A term applied to signal circuit on an overhead or underground line. Line Current: (Power) A current in a straight infinitely long conductor concentrated at the conductor center. Transmission line currents are often treated as line currents for the purpose of calculating the magnetic field away from the conductors. The magnetic field, B, at a distance, R, from a single line current, I, is: B=2I/R, if B is expressed in milligauss, I in ampere, and R in meter. 18-49

Glossary

Linear Materials: If a materials susceptibility (its ability to be magnetized) and permeability (its ability to concentrate flux) are constant over a wide range of magnetic field, the material is said to be a linear material. Linearly Polarized: An electric or magnetic field at a given frequency is linearly polarized when the field vector oscillates in time without changing direction. All the space components are in phase with each other. A single high voltage electrode produces a linearly polarized electric field. A single line current produces a linearly polarized magnetic field. Three-phase electrical installations produce fields that are, in general, not linearly polarized. (See also Polarization). Lineside Signals (UK): Visual signals which are located along the route of a railway (by the side of the line) for observation by the driver. They differ from cab signals that are transmitted from the track to the train to provide visuals on the drivers control desk. Live-Line Maintenance: Maintenance on energized power lines using live-line techniques (with special tools and insulated hot sticks) or bare-hand methods where the worker is directly attached to the conductor and not attached to ground. Live Rail: Synonymous with conductor rail. Local Source: A local field source is located close enough to a shielding enclosure (or region of interest) for its field to be significantly non-uniform over the spatial region of interest. Lock: (Railroad). Electric: A device to prevent or restrict the movement of a lever, a switch, or a movable bridge, unless an electrical device, such as an electromagnet, solenoid, or motor, withdraws the locking member. (See also Electric Locking). Electric (Movable Bridge): An electric lock used in connection with a movable bridge to prevent its operation until released. Electric Switch An electric lock connected with a switch or switch movement to prevent its operation until released. Facing Point: A mechanical lock for a switch, derail or movable point frog, comprising a plunger stand and a plunger that engages a lock rod attached to the switch point to lock the operated unit. Forced Drop: An electric lock in which the locking member is mechanically forced down to the locked position. Movable Bridge: A device used to insure that a movable bridge is in proper position for the movement of trains. Rail (Movable Bridge): A mechanical device used to insure that the movable bridge rails are in proper position for the movement of trains. 18-50

Glossary

Lock and Block: (Historical) A term commonly used for the controlled manual block system. Lock Bar: (Railroad) See Lock Rod. Lock Rod: (Railroad) A rod, attached to the front rod or lug of a switch, movable point frog or derail, through which a locking plunger may extend when the switch points or derail are in the normal or reverse position. Locking: (Railroad) (Historical) (See also Electric Locking, Movable Bridge Locking, Sectional Release, Switch Lever Lock). Electric Switch Lever Locking: (Historical) A general term for route or section locking. Latch-Operated: (Historical) The mechanical locking of an interlocking machine which is actuated by means of the lever latch Lever-Operated: (Historical) The mechanical locking of an interlocking machine that is actuated by means of the lever. Mechanical: (Historical) An arrangement of locking bars, dogs, tappets, cross-locking and other apparatus by means of which the interlocking is effected between the levers of an interlocking machine and so interconnected that their movements must succeed each other in a predetermined order. Preliminary: (Historical) Mechanical locking so arranged that the locking of the lever to prevent it from being moved in conflict with another lever, which is about to be moved, is fully effected before the second lever begins to perform its function. Longitudinal Electric Field: (Historical) Voltage per unit length of a circuit, induced by the magnetic field, when the circuit is in the vicinity of a power line. Longitudinal Electromotive Force has replaced this term. Longitudinal Voltage (Common Mode): The voltage common to all conductors of a group as measured between that group at a given location and an arbitrary reference (usually earth), but not between conductors at each individual end. Lorentz Force: The sideways force imparted to an electric current carrying conductor placed in a magnetic field:

F = I dlB
The equation describing this force (sometimes called the Lorentz relation) can be expanded to include the combined effect (resultant force) of both a magnetic field (on a moving charge) and an electric field (on a charge distribution).

18-51

Glossary

Lossy Shields: A shield that reduces the magnetic field due to absorption and/or reflection attenuation (as opposed to ducting) is called a lossy shield. (See also Absorption, Ducting and Reflection). Low Reactance Phasing: A term that describes multiple electrical circuits arranged in a manner so that the phase sequence for one circuit is placed adjacent to the opposite phase sequence for the other circuit. For example, for a double circuit vertical configuration line, one circuit may have A-B-C phases arranged top to bottom and the adjacent circuit would have C-B-A arranged top to bottom. This method produces lower fields at ground level but somewhat higher conductor surface gradients with the potential for more corona related phenomena. Low reactance phasing is also called reverse phasing, unlike phasing, and cross-phasing. It can be applied to double or multi-circuit transmission or distribution lines, or transmission lines with a lower voltage underbuild. For magnetic field reduction, the method works best when current flow in the adjacent circuits is equal in magnitude and direction. Any deviation from these conditions will lessen the effectiveness of applying low reactance phasing as a field management technique. If the current flow in adjacent circuits is in the opposite direction, the phasing on adjacent circuits should be made similar instead of opposite. (See also Super Bundle Phasing). LRV: Light Rail Vehicle. Lunar White: One of the standard colors used in railroad signaling. Machine: (Centralized Traffic Control) (Railroad) (Historical) A control machine for operation of a specific type of traffic control system of signals and switches. (See also Interlocking Machine). Machine: (Railroad) (See Interlocking, Switch). Interlocking: An assemblage of manually operated levers or equivalent devices, for the control of signals, switches or other units, and including mechanical or circuit locking or both to establish proper sequence of movements. Switch: (See Yard Switch Machine). Machine Cabinet: (Historical) A protection for an interlocking machine. Machine Frame: (Historical) The support for the units of an interlocking machine. Machine Ouadrant: (Historical) The part of an interlocking machine with which the latch block engages. Made Ground: An intentional connection between an electrical conductor and ground. Magnetic Circuit: The complete path of magnetic lines of force. A magnetic circuit can be compared to an electric circuit so that the magnetic flux corresponds to electric current, magnetomotive force corresponds to electromotive force, and reluctance is comparable to resistance. (See also Reluctance). 18-52

Glossary

Magnetic Dipole: (See Dipole). Magnetic Field: Magnetic field is usually referred to as the vector field of magnetic flux density. (See Magnetic Flux Density). Magnetic Field Strength: A magnetic field exists in the region near a permanent magnet or current-carrying conductor and can exert a force on electric currents (or an independent pole) placed in them. The magnetic field strength (H) at a point in space is a vector defined by its space components along three orthogonal axes. For steady-state sinusoidal fields, each space component is a complex number or phasor. The magnitudes of the components are expressed by their rms values of magnetomotive force (mmf) per unit length. In a multi-phase environment, such as near a threephase electric power line, the field is characterized by a vector rotating in a plane where it describes an ellipse whose semi-major axis represents the magnitude and direction of the maximum value of the magnetic field, and whose semi-minor axis represents the magnitude and direction of the field a quarter cycle later at its minimum value. The unit of magnetic field strength in the MKS (SI) system is the ampere-turn per meter, or simply, the ampere per meter (A/m), which is mmf per unit length. One ampere-turn per meter is the magnetic field strength in the interior of an elongated, uniformly wound solenoid that is excited with a linear current density in its winding of one ampere per meter -3 of axial distance. (In the CGS system, the oersted is the unit of magnetic field strength: 4 x 10 oersteds = 1 A/m). Sometimes the magnetic field strength (H) is described (incorrectly) using the units of its magnetic flux density (B) with MKS (SI) units of tesla (or T) and CGS units of gauss (or mG). One milligauss equals 0.1 microtesla. Magnetic field and magnetic flux density are related by the permeability of the material or medium in which they are characterized. The ratio of flux density (B) to field strength (H) is the permeability. (See also Magnetic Flux Density and Permeability). Magnetic Flux: Magnetic Flux is the integral of the normal component of the magnetic flux density over a surface. Magnetic Flux Density: The number of magnetic flux (force) lines per unit of cross-sectional area that permeates a magnetic field is the magnetic flux density (B). The path of an independent (or isolated) pole in a magnetic field suggests a line of flux. A line of flux is a line so drawn that a tangent to it at any point indicates the direction of the magnetic field. The unit of magnetic flux () in the MKS (SI) system is the weber (Wb); in the CGS system, the unit of magnetic flux is 8 the maxwell (Mx). One maxwell equals one magnetic flux line, and one weber equals 1 x 10 lines. The lines of flux () perpendicular to a specific area (A) in the magnetic field are collectively called the magnetic flux density. Therefore, the magnetic flux density, B, is given by: B = /A. According to Maxwells equations the net magnetic flux through any closed surface is zero. Magnetic flux density in the MKS (SI) system is expressed in webers per square meter (Wb/m2), or tesla (T); in the CGS system, magnetic flux density is expressed in maxwell per square centimeter (Mx/cm2) or gauss (G). These units have the following equivalence: One tesla equals 104 gauss, and 0.1T = 1mG. The magnetic field strength (H) is related to flux density (B) by the permeability of the medium. (In air, 4 mG equals 1 A/m). (See also Flux, Gauss, and Tesla).

18-53

Glossary

Magnetic Materials: Materials that are readily magnetized in a magnetic field are called magnetic materials. When one considers materials simply as either magnetic or nonmagnetic, this division is really based on the strong magnetic properties of iron. For example, iron, nickel and cobalt are common examples of magnetic materials and air, paper, and wood are examples of nonmagnetic materials. However, weak magnetic materials (sometimes called nonmagnetic materials) can be important in some applications. For this reason, a more exact classification includes the following three groups: Ferromagnetic materials. These include iron, steel, nickel, cobalt, and commercial alloys such as alnico and Permalloy. They become strongly magnetized, in the same direction as the magnetizing field, with high values of permeability from 50 to 50,000. Permalloy has a relative permeability (with respect to air of 1) of 100,000, but is easily saturated at relatively low values of flux density. Paramagnetic materials. These include aluminum, platinum, manganese, and chromium. Their permeability is slightly more than 1. They become very weakly magnetized in the same direction as the magnetizing field. Diamagnetic materials. These include bismuth, antimony, copper, zinc, mercury, gold, and silver. Their permeability is less than 1. They become very weakly magnetized but in the opposite direction from the magnetizing field. Magnetic Pole Strength: The strength, or the attractive (or repulsive) power of a magnetic pole, is measured by the number of unit poles to which each pole is equivalent. The total number of lines of force emanating from a magnetic pole is called the total magnetic flux. The number of flux lines from a pole of unit strength is 4 lines (or 4 maxwells). The number of magnetic lines emanating from a pole of strength m is 4m lines. (Note: 1 maxwell = 1 line, and 1 weber = 108 lines). (See also Flux, Maxwell, and Weber). Magnetic Quadrupole: (See Quadrupole). Magnetism: The property of materials to be magnetized. A material is magnetized when the orientation of its individual magnetic dipole moments is affected by an externally produced magnetic field. There are different types of magnetism: paramagnetism, diamagnetism, and ferromagnetism. Paramagnetic materials, such as aluminum, have individual magnetic dipole moment randomly oriented in the absence of external magnetic field and the material is unmagnetized. When magnetic field is applied the individual magnetic moments align themselves with the field. When the field is removed the magnetic dipoles return to their random orientation. In diamagnetic materials, such as copper, dipole moments are induced in a direction opposite to that of the applied fields. Ferromagnetic materials, such as iron, behave similarly to paramagnetic materials except they become much more magnetized than paramagnetic materials with a given applied field and, if the field is removed; many of the magnetic dipoles remain aligned. Paramagnetism and diamagnetism are extremely weak phenomena compared to ferromagnetism. For most practical considerations they can be completely neglected. It is common practice to refer to ferromagnetic materials as magnetic materials. (See also Magnetic Materials, Ferromagnetism, Diamagnetism, and Paramagnetism).

18-54

Glossary

Magnetization: The degree to which a material is magnetized is quantified by the vector quantity magnetization (M). The magnetization is defined as the net dipole moment per unit volume, or the magnetic dipole density. Just as mass density (kg/m3) quantifies the degree to which matter is concentrated in materials, magnetization, M, quantifies the degree to which dipole moments are concentrated in materials. Its units are those of dipole moment per unit volume (Am2/m3=A/m), or the same as magnetic field, H, in A/m; its direction is that of the preferred direction of the individual atomic dipoles. The degree to which a material becomes magnetized (the value of M) with a given applied field depends on the properties of the material itself. (See also Diamagnetic, Dipole Moment, Ferromagnetic, Paramagnetic and Susceptibility). Magnetization Curve: To illustrate the relation of magnetic flux density (B) to applied magnetic field (H), a graph showing B (ordinate) as a function of H (abscissa) is used. The line or curve showing B as a function of H on such a B-H graph is called a magnetization curve. Because of the complex manner in which B changes with H, it is not possible to express B as a concise mathematical function of H. Instead, the relationship is expressed in the form of tables or graphs for different materials. A magnetization curve shows that the value for a materials permeability changes with H. It should be noted that the materials permeability () is not the slope of the curve, which is given by dB/dH, but is equal to the ratio B/H. (See also Hysteresis). Magnetomotive Force: Ampres law equates the magnetic field integrated around a closed path or loop to the electric current through the loop. The magnetomotive force (mmf) is equal to this enclosed current (I), and it is considered to produce flux in a magnetic material against the materials resistance to the production of magnetic flux (this opposition to flux is called reluctance). In a magnetic circuit, mmf (in ampere-turns) is analogous to electromotive force emf (volts) in an electric circuit. If the path of integration (or loop) encloses a number (N) of turns (coils) of wire, each with a current (I) in the same direction, then mmf is given as NI, or ampere-turns in the MKS (SI) system or gilberts (CGS). One ampere-turn per meter equals 4 x 10-1 gilberts. (See also Ampere-Turn, Electromotive Force, Flux and Reluctance). Magnetostatic: (See Quasistatic). Magnets: The two broad classes of magnets are permanent magnets and electromagnets. An electromagnet needs current from an external source to maintain its magnetic field (unless a superconductor is used). With a permanent magnet, not only is its magnetic field present without any external current, but the magnet (if stored properly) can maintain its strength indefinitely. The basis of all magnetic effects is the magnetic field associated with electric charges in motion. Within the atom, the motion of its orbital electrons generates a magnetic field. There are two kinds of electron motion in the atom. First is the electron revolving in its orbit around the nucleus. This motion provides a diamagnetic effect. However, this magnetic effect is weak because thermal agitation (Johnson Noise) at normal room temperature results in random directions that neutralize each other. More effective is the magnetic effect from the motion of each electron spinning on its own axis. The spinning electron serves as a tiny permanent magnet. Opposite spins provide opposite polarities. Two electrons spinning in opposite directions form a pair, neutralizing the magnetic fields. In the atoms of ferromagnetic materials, however, there are many unpaired electrons with spins in the same direction, resulting in a strong magnetic effect. In terms of molecular structure, iron atoms are grouped in microscopically small arrangements called domains. Each domain is an elementary dipole magnet, with two opposite poles. In crystal form, the iron atoms have domains that are parallel to the axes of the crystal. Still, the domains can point in different directions, because of the different axes. When the material becomes magnetized by an external magnetic field, though, 18-55

Glossary

the domains become aligned in the same direction. With permanent magnetic materials, the alignment remains after the external field is removed. (See also Dipole Moment, Domain, Electromagnet, Ferromagnetic and Permanent Magnet). Main Track: A railroad track (other than an auxiliary track) extending through yards and between stations, upon which trains are operated on a timetable or train order, or both, or the use of which is governed by block signals. Manipulation Chart: (Railroad) (Historical) A statement in tabulated form showing the sequence in which levers or other devices must be operated. Manual Block Signal System: (Railroad) A block or a series of consecutive blocks, governed by block signals operated manually, upon information by telegraph, telephone or other means of communication. Controlled: A series of consecutive blocks governed by block signals, controlled by continuous track circuits, operated manually upon information by telegraph, telephone or other means of communication, and so constructed as to require the cooperation of the signalmen at both ends of the block to display a Clear or a Permissive block signal. Marker Light: (Railroad) A light which by its color or position, or both, qualifies the signal aspect. Master Controller: Train drivers power control device located in the cab. Master Unit (Control Office): (Railroad) A group of relays and circuits in a code system that generates the control codes and receives and interprets the indication codes. Maximum Continuous Operating Voltage (MCOV): The maximum designated rms value of power frequency that may be applied continuously between the terminals of the surge protective device without degradation of the device. Maximum Field: The space component of a power frequency field (electric or magnetic) in the direction where the field is maximum. The maximum electric field is measured by orienting a free-body electric field meter in various ways until the maximum reading is obtained. The maximum magnetic field is measured by orienting a single axis sensor, e.g. a coil, in various ways until the maximum reading is obtained. The measurement of three orthogonal field components is not sufficient for the calculation of the maximum field. The square root of the sum of the squares of three orthogonal field components is the field resultant. The field resultant coincides with the maximum field, only when the field is linearly polarized. Power frequency fields are often polarized. In these cases the field resultant is greater than the maximum field, by a factor that can be as large as 2 . Maxwell: The unit of magnetic flux is the maxwell in the CGS system. One maxwell (Mx) equals one magnetic line of force (or line of induction). In the MKS (SI) system the unit of magnetic flux is the weber. One weber (Wb) equals 1x108 lines (or Mx). Since the weber is a large unit for typical fields, the microweber unit can be used (1Wb = 100 lines or 100 Mx).

18-56

Glossary

Maxwells Equations: James Clerk Maxwell (1831-1879) put the laws of electromagnetics in the form we know them today. Maxwells equations are used constantly and universally in the solution of a wide variety of practical problems. The scope of Maxwells equations is remarkable, including the fundamental principles of electromagnetic and optical devices. The relations stated by Maxwell, known as Maxwells Equations, consist of four expressions: one derived from Ampres Law, one from Faradays Law, and two derived from Gauss Law. Maxwells equations are given here in general form and for free space. Maxwells Equations: General Set
Integral Form Amperes Law (Extended by Maxwell to describe enclosed current)
D

Point Form

H dl = S (J c + t ) dS
Faradays Law of Induction

H = Jc +

D t

(Describes electric effect of a changing magnetic field)

E dl = -S t dS
Gauss Law for Electricity

E = -

B t

(Describes relation between electric flux and charge)

S D dS = v dv
Gauss Law for Magnetism

D=

(Describes the nonexistence of a magnetic monopole)

S B dS = 0

B = 0

Note: Jc = Conduction current density and = charge density, = Closed path integral, S = Surface integral, S = Closed surface integral, v = Volume integral. For the special case of free space, where the conduction current density, Jc , and the charge density, , are zero, the equations reduce to the following simpler Free-Space form:

18-57

Glossary

Maxwells Equations: Free-Space Set


Integral Form Amperes Law (Extended by Maxwell to describe enclosed current)
D D t

Point Form

H dl = S t dS
Faradays Law of Induction

H =

(Describes electric effect of a changing magnetic field)

E dl = -S t dS
Gauss Law for Electricity

E = -

B t

(Describes relation between electric flux and charge)

s D dS = 0
Gauss Law for Magnetism

D = 0

(Describes the nonexistence of a magnetic monopole)

S B dS = 0

B = 0

Milligauss: A unit in the CGS Systems used to describe magnetic flux density (flux lines per unit of cross-sectional area) is the milligauss. It is one thousandth of a gauss (0.001G); it has the following equivalence with the tesla unit of the MKS (SI) system: 1mG = 0.1T. See also Gauss, Magnetic Flux Density, and Tesla. Motion Sensitive Device: A device used to sense the presence, motion and direction of travel of a train. A device used to detect the movement of a train. MMF: (See Magnetomotive Force). MPF: (Railroad) Movable Point Frog. Multipoint Ground: (See Multiple Point Grounding). Multiple Shields: For some shielding applications, a single shield may saturate in a high flux density field. Rather than make the shield excessively thick, the solution could be to use multiple (nested) shields of thinner material separated by air gaps (and preferably not connected). Sometimes, the outer shield is of a material with lower permeability that is not easily saturated and the inner shield is of a higher permeability material. Multiple shielding layers can provide a greater shielding effectiveness than a single shield. Because of successive reflections in the space between 18-58

Glossary

two shields, the total shielding effectiveness is less than the sum for the separate shields. (For example, in the highly effective portion of the frequency spectrum, practical experience indicates that two 70-dB shields may result in approximately 120-dB shielding effectiveness.) (See also Proximity Factor). Multiple Point Grounding: A term used to describe multiple connections to earth by use of driven grounding rods, attachments to buried metallic pipes, or other grounding methods. Multiple-Way Switches: (Power) Switches that can connect a hot wire to either one of two or more other wires for the purpose of switching electrical devices, usually light fixtures, on and off independently from two or more locations. The wiring of multiple-way switches ordinarily does not produce magnetic fields. However, they can be connected in ways that create significant magnetic fields when the light fixtures are switched on. Multipoles: Any general current distribution can be represented mathematically by a binomial expansion with an infinite number of terms. The first four terms in the series are called the monopole, dipole, quadrupole, and octopole terms, respectively. Collectively, they are referred to as multipoles. Mumetal: The generic term for all high nickel content, very high permeability nickel-iron alloys is mumetal (See ASTM A753-85 and MIL-N-14411C). MuMetal (with two capital Ms) is the trademark for a 75% nickel-5% copper shielding alloy made by Allegheny Ludlum, Inc. Similar high permeability ferromagnetic materials are Hipernom (80% nickel, balance iron) and Hipernik (50% nickel, balance iron) made by Carpenter Technology. All of the very high permeability materials ( of tens to hundreds of thousands) are excellent shielding materials because of their ability to readily provide a low reluctance path for magnetic flux lines. The permeability of a material increases with magnetic flux density, so that attenuation (shielding) generally increases as the magnetic field becomes stronger (and the material is more strongly magnetized). The limiting point is the maximum permeability of the material (the knee of the magnetization curve) - as flux density increases beyond the maximum point; the materials permeability decreases rapidly to saturation. Permeability also decreases with higher frequency. High permeability materials such as mumetal will suffer a significant loss in permeability due to fabrication processes (welding and mechanical work) and must be specially heat-treated (hydrogen annealed) under controlled conditions to restore and/or increase permeability of the material to its optimum value. Shielding enclosures fabricated from high-permeability materials such as mumetal are increasingly sensitive to strain as their permeability rises and repeated dropping or mechanical shock can destroy shielding effectiveness by reducing permeability. Therefore, mumetal shields should be carefully handled after fabrication to avoid loss of permeability. (One leading shielding manufacturer has advised that they be treated like glass.) (See also Ferromagnetism, Flux, Permeability, and Reluctance). Mutcd: (Railroad) Manual on Uniform Traffic Control Devices. National Electrical Code (NEC): The National Electrical Code (NEC) is a published volume of rules whose purpose is the practical safeguarding of persons and property from hazards arising from the use of electricity. The NEC states that its rules cover installations of electric conductors and equipment that connect to the supply of electricity, including conductors and equipment within or on public and private buildings or other structures, including mobile homes, 18-59

Glossary

recreational vehicles, floating buildings, and other outside premises such as yards, carnival, parking and other lots and industrial substations, installations of optical fiber cable, and installations in buildings used by the electric company (offices, warehouses, garages, machines shops, etc.) The original code was developed in 1897 as a result of the efforts of various insurance, electrical, architectural and related interests. The National Fire Protection Association has sponsored the present NEC for 87 years. The NEC is updated and revised, as necessary, on a 3-year review cycle. The NEC committee includes representatives from insurance firms, electric utilities, Underwriters Laboratories, OSHA, organized labor, electrical contractors, telecommunications companies, equipment manufacturers, academia, heavy industry, chemical companies, regulatory and inspection organizations, engineering firms, water supply, gas and oil industry, electrical testing firms, cable and wire companies, electronics firms, computer companies, U.S. government agencies, etc. The NEC applies starting at the service point (the electric service meter) and includes facilities within the customers installation. The NEC does not cover the electric supply system - that is covered by the National Electrical Safety Code. National Electrical Safety Code (NESC): The National Electrical Safety Code (NESC) is a published volume of rules whose purpose is the practical safeguarding of persons during the installation, operation, or maintenance of electric supply and communication lines and associated equipment. The NESC states that its rules contain the basic provisions that are considered necessary for the safety of employees and the public and the NESC is not intended as a design specification or instruction manual. Work on the NESC started 85 years ago at the National Bureau of Standards. The NESC is applied to the electric supply system up to the electric service point of a customer; it does not cover premise wiring. The National Electrical Code (NEC) covers wiring from the electric service or supply point (the electric service meter). The present NESC is an American National Standard (ANSIC2-1997) with IEEE as the administrative secretariat. The NESC is revised, as necessary, on a 5-year review cycle. The NESC is a consensus document prepared by those substantially concerned with its scope and provisions. The standards committee membership includes representatives, among others, from the electric power industry, telephone industry, insurance companies, railroads, organized labor, regulatory agencies, contractors, equipment manufacturers, cable TV, and safety council. In some states in the U.S., other codes govern subjects covered by the NESC (e.g. For overhead transmission lines: General Order No. 95 in California and General Order No. 6 in Hawaii). NBS: (United States) National Bureau of Standards, now called NIST National Institute Standards and Technology. Near Field: The near field is a term used to denote the region relatively close to a radio frequency (RF) antenna compared to the wavelength corresponding to the frequency of operation. In this region the relationship between the electric and magnetic fields is complex and not fixed as in the far field, and in which the power density does not necessarily decrease inversely with the square of the distance. This region is sometimes defined as closer than about one-sixth of the wavelength. In the near field region the electric and magnetic fields can be determined, independently of each other, from the free-charge distribution and the free-current distribution respectively. The spatial variability of the near field can be large. The near field predominately contains reactive energy that enters space but returns to the antenna (this is different from energy that is radiated away from the antenna and propagates through space). Negative Sequence Component: (See Symmetrical Components). 18-60

Glossary

NEC: (See National Electrical Code). NEMA: National Electrical Manufacturers Association. NESC: (See National Electrical Safety Code). Nested Shields: (See Multiple Shields). Net Current: (Power) When the instantaneous values of all the current carrying conductors (including neutral conductors) in a circuit do not sum to zero then a net current is said to exist. Net currents are important in field management principles because they attenuate very slowly (with the inverse of distance) and are difficult to cancel with methods commonly used for balance phase currents. Net currents must always return to the source transformer. Net currents may flow in the neutrals of other circuits, in the earth, or along any other conductive path to return to their source transformer. Net Load Current: (Power) The electrical service of US residences is generally provided by two hot wires and a neutral. The voltages between the two hot wires and the neutral have opposite phases and so, approximately, have the two load currents. The vectorial sum of the two load currents is called the net load current. This is the current that is measured by placing a clamp-on ammeter around the two hot wires (not the neutral). Neutral Bus (or Bar): (Power) The neutral bus of an electrical panel is the electrode to which all the neutral conductors of the electrical circuits of the panel are connected. The neutral bus of the service entrance panel (main panel) is connected to the panel case and to the grounding electrode. The neutral bus of a sub-panel is isolated from the panel case. Neutral Conductor: (Power) A conductor in an electric circuit designed to carry the unbalanced current from the other conductors. In wye-connected transformers, the neutral may or may not be grounded and it provides a return path for unbalanced phase current. In delta-connected transformers, a neural conductor is not required. Neutral Shared: (Power) A shared neutral is a neutral conductor that carries the return current for more than one circuit in an electrical system. In a shared neutral arrangement, there are two or more electrical circuits with individual phase wires but only one shared neutral wire. If the phase conductors are not close to the neutral, a current loop is created and this can cause an increase magnetic field level. Failure to keep the neutral close to all the associated phase conducts is a violation of the National Electrical Code Section 300-3 (b). NIST: The National Institute of Standards and Technology (NIST), formerly the National Bureau of Standards (NBS), was established by Congress in 1901 to support industry, commerce, scientific institutions, and all branches of Government. For nearly 100 years the NIST/NBS laboratories have worked with industry and government to advance measurement science and develop standards. NIST maintains measurement and standards laboratories that provide technical leadership for vital components of the nations technology infrastructure. EMF calibration facilities generally utilize precision components or methods that are traceable to NIST for their accuracy.

18-61

Glossary

Non-Vital-Circuit: Any circuit the function of which does not affect the safety of train operation. Nonlinear Materials: If a materials susceptibility (its ability to be magnetized) is a constant (not a function of magnetic field, H) the material is said to be linear; and if the susceptibility is not constant, is said to be a nonlinear material. Ferromagnetic materials are known to be nonlinear (susceptibility and permeability are functions of H). Further complicating the issue is the facts that not only do susceptibility and permeability depend on the value of H, but also on the past magnetic history of the material because of a property called hysteresis. (See also Hysteresis). Nonmagnetic Materials: Materials that essentially are not magnetized (not affected) by a magnetic field are generally called nonmagnetic materials. Examples of nonmagnetic materials include air, paper, wood, and plastics. Weak magnetic materials (diamagnetic and paramagnetic) are sometimes referred to as nonmagnetic materials. (See also Magnetic Materials). Normal Clear: (or Stop) (Railroad) (See Automatic Block Signal System). Normal Mode: See Differential Mode Voltage. Normal Position: (Railroad) The position in which signal and other devices are assumed to normally lie, according to rule, convention or otherwise, i.e., Stop aspect displayed, switch set for main track, devices energized or de-energized, etc. (See also Reverse Position). Number Plate: (Railroad) A device fastened to signal apparatus for the purpose of identification. Occupied Switch (OS): (Railroad) This is the track circuit that detects the presence of a train in the immediate vicinity of a power-operated track switch, and prevents the switch from changing positions while a train is occupying that circuit. Office Blocking: (Sometimes called Machine Blocking). OHSW: Overhead Shield Wires (see Shield Wires). Open Wire Line (Railroad) An overhead wire line consisting of single conductors as opposed to multiple-conductor cables. Operated Unit: (Railroad) A switch, signal, lock or other device, which it is the function of a lever or other operating means to operate. Operator: (Railroad) (See Control Operator). Opposing Signals: Railroad signals that govern movements in opposite directions on the same track. Opposing Train: A train, the movement of which is in a direction opposite to and toward another train on the same track. 18-62

Glossary

Oersted: The unit of magnetic field strength in the CGS electromagnetic system. The oersted is defined as the magnetic field strength in the interior of an elongated, uniformly wound solenoid that is excited with a linear current density in its winding of one abampere (10 amperes) per 4 centimeters of axial length. (In air, a 1-oersted magnetic field produces 1 gauss of magnetic flux density). The oersted has the following equivalence with the MKS (SI) unit of magnetic field strength: 4 x 10-3 Oe = 1 A/m. OS: Occupied switch. An alternate historical meaning for OS is On Sheet. In this meaning, as a train passes, the operator would report the time to the dispatcher to record on the Train Sheet which was maintained at the dispatchers desk. This information is often retained for some years as a record and could then be used whenever later conflicts arose or rule violation incidents were being investigated. Outlying Switch: (Railroad) A switch not included in a nearby interlocking or controlled point. It is not necessarily under the control of the operator or dispatcher. Overhead: (Railroad) A generic term (as in the overhead) referring to electric traction supply wires suspended over the track for current collection by trains. Also known as overhead line, OLE (overhead line equipment), or catenary after the line suspension system. A pantograph collects current on the roof of the train or locomotive. Overlap: (Railroad). Breaking Distance: The safe braking distance beyond a signal provided in case the train fails to stop at the signal when it is showing a danger aspect. Grade Crossing Approach: The portion of the track used by two separate highway grade crossings. Signal Control Distance: The distance the control of one signal extends into the territory that another signal, or signals, governs. Overlap Block Signal System: (Railroad) A block signal system in which the control of a signal, or signals, extends into the territory that another signal, or signals, governs, so that one or more opposing signals display an aspect indicating Stop. Overrun: (Railroad) Distance allowed beyond a normal stopping point in case a train fails to stop in the correct position. The distance is dependent upon speed and braking capacity of the train. P-Wire Control: An analogue form of electro-pneumatic brake control using a single wire carrying a pulse width modulated signal. The current level on the circuit determines the rate of brake demanded. Zero current initiates an emergency brake demand. This type of control can eliminate the need for a brake pipe, a source of much trouble on trains.

18-63

Glossary

Pantograph: A folding traction current collection device mounted on the roof of a vehicle on a railway employing an overhead electric supply system. Modern pantographs are sophisticated aero-dynamically designed devices which can operate at high speeds without loss of contact and with built-in safety devices which reduce the risk of damage to wires in the event of a fault. A common problem is when a pantograph catches above the wire and pulls it down for considerable distances before the crew notices it and the train stopped. Modern pantographs are fitted with automatic detection and lowering devices. The horns (curved edges) of the pantograph are equipped with frangible pneumatic sensors that, if broken by a wire support, cause the detector system to lower the pantograph. Parallel Plates: A method used to calibrate free-body electric field meters involves placing the meter between a pair of parallel conductive plates. The parallel plates should have a specific geometry so that there is a 2:1 ratio between plate dimensions and the spacing between them. Typical parallel plates are 1.5-meter squares with spacing between the plates of 0.75 meters. A variable voltage source is used to apply a potential across the plates creating a relatively uniform vertical electric field in the central region between the plates. The meter to be calibrated is placed between the plates using the insulating handle that is supplied with this type of instrument. The method requires accurate dimensions for the plates and careful measurement of the applied voltage by use of a precision component or method traceable to NIST to guarantee accuracy. Paramagnetism: Most materials show some magnetic effects due to electric charges in motion at the molecular, atomic, and subatomic levels (e.g. electrons revolve around the nucleus of an atom and also spin about their own axis as they move as paired or unpaired electrons). With the exception of the ferromagnetic group of materials, these magnetic effects are weak. Depending on their magnetic behavior, materials can be classified as diamagnetic, paramagnetic, or ferromagnetic. In paramagnetic materials, a magnetic moment is induced into the materials atomic structure by an applied magnetic field in the same direction as that of the applied field. Aluminum, platinum, and palladium are examples of materials that are very feebly attracted by a strong applied magnetic field, and are said to be paramagnetic. This type of behavior is called paramagnetism. Passive Loops: Passive loops is a term used to describe a system of conductors arranged in a path (loop) near to a transmission or distribution line, or other source, that uses induced currents to reduce the source magnetic field within a region of interest. The magnetic field due to the power line (or other source) induces currents in the conductive loops that, in turn, create another magnetic field that partially cancels the original magnetic field. (See also Cancellation Loops). Passive Shielding: Magnetic fields can be reduced (shielded) by establishing currents in wires such that the fields produced by those currents oppose the fields to be reduced. Passive shielding (or shielding with passive conductors) refers to use of currents induced in conductors by existing (or ambient) magnetic fields to reduce (shield) these fields in a certain region. Penetration Loss: (See Absorption Loss). Permanent Magnets: Permanent Magnets are made of hard (not easily magnetized) magnetic materials, such as cobalt steel, magnetized by induction in the manufacturing process. A very strong field is needed for induction in these materials. When the magnetizing field is removed, however, a residual induction (retentivity) makes the material a permanent magnet. A common permanent 18-64

Glossary

magnet material is alnico, a commercial alloy of aluminum, nickel, and iron, with cobalt, copper, and titanium added to produce a number of grades. Commercial permanent magnets (when stored properly with a keeper) will last indefinitely if not subjected to high temperatures, to physical shock, or to a strong demagnetizing field. If the magnet becomes hot, however, the molecular structure can be rearranged, resulting in loss of magnetism that is not recovered after cooling. The point at which a magnetic material loses its ferromagnetic properties is the Curie temperature. For iron, this temperature is about 770C, when the relative permeability drops to unity (the same as air). A permanent magnet does not become exhausted with use, as its magnetic properties are determined by the structure of the internal atoms and molecules. (See also Curie Temperature). Permanent Way: A generic term for railway track, referring to the rails, ties (sleepers UK) and ballast. The term permanent arose to distinguish it from the temporary track laid during the construction of the railway. Permeability: The property of a material by which it changes the flux density in a magnetic field from the value in air is called its permeability (it is analogous to permittivity, , for electric field). Permeability, , is a physical constant (for a given temperature and material density) that reflects a material or mediums ability to concentrate magnetic flux lines. It is a function of the magnetic properties of the material (or medium). The permeability of a vacuum, , is equal to 4 x 10-7 -7 weber/amp-meter in the MKS (SI) system (it is 4 x 10 henry/meter in the CGS system). The permeability of a vacuum and air are very nearly the same (and usually are taken as such). The permeability of diamagnetic materials are slightly less than , while paramagnetic materials slightly more. Ferromagnetic materials have very large values of permeability. Permeability is the ratio of magnetic flux density (B) to magnetic field (H) in the medium; so that for free space:
=
B H

and, B = H . Therefore, in free space, the conversion between magnetic field (H=A/m)

and magnetic flux density (B=gauss or tesla) is as follows: 1mG = 0.1 T = (1/4) A/m. In materials or mediums other than free space, it may be necessary to consider the magnetization (M) of the material such that:
=
B H +M

and, B = (H + M) . The permeability of a material is also a measure of its ability to be

magnetized and is defined with respect to the permeability of free space ( ) and the susceptibility ( ) of the material as follows: = (1 + ). For ferromagnetic materials, permeability is a nonlinear property. Very frequently in engineering applications, the dimensionless relative permeability (r) is used to quantify a materials permeability () with respect to free space, as follows:
r = . Permeability is a function of (among other things), flux density, frequency, and material

thickness. (See also Curie Temperature, Magnetic field, Magnetic Flux Density, Magnetization, Nonlinear materials, Relative Permeability, and Susceptibility). Permissive: (Railroad) (See Block). 18-65

Glossary

Permittivity: The property of a material by which it changes the electric flux density in an electric field from the value in free space (a vacuum) is called permittivity (it is analogous to permeability, , for magnetic field). Permittivity, , is a physical constant (for a given temperature and material density) that reflects a material or mediums ability to concentrate -12 electric flux lines. The permittivity of a vacuum, D , is equal to 8.854 x 10 farad/meter. Permittivity is the ratio of electric flux density (D) to the electric field (E) in the medium; so that for a vacuum:

D and, D = D E . E

Phantom Aspect: (Railroad) An aspect displayed by a light signal, different from the aspect intended, caused by a light from an external source being reflected by the optical system of the signal. Phase: (Power) The phase or phase conductor is any conductor (other than the neutral conductor) in an electric power circuit designed to be energized at the nominal system voltage and carry power. In a three-phase power system there are three phases consisting of one or more conductors per phase. For a balanced three phase electrical system, the phase conductors all have the same magnitude of current and are separated by 120 electrical degrees. Phase Difference: The time in electrical degrees by which one electrical phase leads or lags another. Phasing (Of Power Lines): The order in which different phases of multi-phase lines are placed (e.g. two three-phase lines on the same structure, several three-phase cable systems in the same ducts). For double-circuit lines, for instance, there are several possible phasing, such as superbundle (or same phasing) and low reactance (or reverse phasing). Phasor: A quantity with a sinusoidal time variation described by a magnitude and an angle (phase angle) in time. Q = q sin ( + ) . Pick Path: (Railroad) The portion of a circuit used to initially energize a relay coil. Pick-Up Value: (Railroad) (See also Working Value) The electrical value, when applied to an electromagnetic instrument, that will cause the moving component to move to the position that will just close the front contacts or visually indicate its energized position. Pipe-Type Cable: A pressure cable in which the container for the electric conductors and pressurized medium is a rigid metal pipe. The pipe-type system is pressurized, usually with a dielectric liquid, and is commonly called a high-pressure-fluid-filled (HPFF) cable system. This cable system was called high-pressure-oil filled (HPOF) until the 1970s, when synthetic dielectric liquids began to replace the mineral oils that had been used earlier. Pipe-type cables rated up to 138kV may be pressurized with nitrogen, and then are called high-pressure-gas-filled (HPGF) cables. In general a cable simply called pipe-type refers to an HPFF cable. The first pipe-type cable system was a 66kV circuit installed in 1932 by Philadelphia Electric Company. Over 80% of the transmission cable in the United States is HPFF. An excellent reference is the EPRI Underground Transmission System Reference Book. PLC: (See Power Line Carrier). 18-66

Glossary

Point Source: A source of magnetic field that can be equated to a magnetic dipole (or, rarely, to a higher order element) concentrated at a point. Several operator sources may be considered point sources. A point source is characterized by a center and by the magnitude and direction of its dipole moment. The magnetic field decays with the third power of the distance from the center of a point source. Points (UK): Referred to in North America as a switch. The track work mechanism where a track divides into two. The rails are specially shaped to allow a smooth transition from the main track to the diverging track. Also referred to as a turnout. Point Locks (UK): Mechanical devices attached to points (switches in North America) to ensure that they remain fixed for the passage of a train through them. In many countries, they are a legal requirement where passenger trains are operated. Points are also electrically locked by their control system and by track circuits occupied by a train passing through them. Polarity: A term used to describe an electrical state or condition of a quantity (e.g. charge, current, field) with respect to a reference point or condition. For time-invariant (or dc) quantities, the extent to which something is positive or negative describes its polarity. For time-varying (or ac) quantities, the extent to which things have the same instantaneous angular variation with time describes similar polarity. For example, relative polarity of a transformer is a designation of the relative instantaneous direction of current in its leads (the same principle applies to the polarity of all transformer windings and here relative polarity can be either additive or subtractive). Polarization: The description of the angular variation with time of either the electric or the magnetic field vector at a fixed point is the field polarization. If the spatial components of the field are all in phase with each other, the field is said to be linearly polarized (the field vector will oscillate back and forth along a line with time). If, on the other hand, the individual spatial components of the field are not all in phase, the field will be elliptically polarized (the field vector traces out an ellipse with time). An elliptically polarized field can be described by its semi-major axis (or maximum value) and its semi-minor axis (minimum value) a quarter cycle later. The field ellipse is completely specified by its degree of polarization (ratio of minor/major axis) and its orientation (the spatial direction of its major axis). A polarized field with a degree of polarization of unity (major and minor axes are equal) is said to be circularly polarized since its field ellipse will trace out a circle with time. Polarized Circuit: A path in which the direction of flow of an electric current is reversed under certain conditions. Pole: (Power) An isolated unit of either of two opposed or differentiated sources of fundamental electric charge or magnetism is called a pole. (See also Dipole and Monopole). Pole Changer: A device by which the direction of current flow in an electrical circuit may be changed. Poles: (Power) Electric power line poles are usually made of treated wood, galvanized or painted steel, or concrete.

18-67

Glossary

Positive Sequence Component: See Symmetrical Components. Possession (UK): (Railroad) When a section of track is required for maintenance and trains cannot run it is handed over by the operators to the engineers, who take possession. Special protective measures are used to prevent access by unauthorized trains. When the track is returned to the operators, the engineers give up possession. Potential: Potentials are functions in space from which vector fields can be obtained by taking spatial derivatives. For instance, the electric field (E) can be obtained by taking the negative gradient of the electric potential (V): E = V . This expression defines the potential V, i.e. the electric potential is the scalar function such that its negative gradient is the electric field. Similarly, there is a potential for magnetic fields. However, this potential is a vector function, A, and is defined by: B = A, and is called the vector potential for the magnetic vector potential. For the special case of regions of space where there are no currents, there is also a magnetic scalar potential, U, such that: B = U. Potential False Proceed Condition (PFPC): (Historical) A condition existing in railroad signal systems, devices, or appliances, when no train is present, under which a false proceed failure would have occurred had a locomotive or train approached or entered a section of track occupied by another train. Potential Point: (Power) A point at which the potential is calculated for the solution of electric field problems using the charge simulation method. (See also Charge Simulation Method). Power Factor: (Power) The ratio of power (watts or kilowatts) to the product of voltage and current (volt-amperes or kilovolt-amperes). Also called the ratio of real power to apparent power. Power Line Calculator: The Power Line Calculator is an EPRI computer program used to analyze current and phase angle variables, among others, for transmission and distribution lines. The Power Line Calculator is available as a stand-alone Microsoft Windows application. The software is also integrated into the EMF Workstation and other EPRI software (SUBCALC and RESICALC) to facilitate analysis of power line parameters. The most current version of the Power Line Calculator is 2.0. The Power Line Calculator is based on the method of symmetrical components. The software allows the current in an electrical circuit to be specified in terms of apparent, real, and reactive power, the phase currents and their time angles, or as a set of symmetrical components. The software provides 18 parameters that can be used to specify the currents on a power line. Unbalanced conditions can be evaluated as well. For underdetermined systems, basic defaults using reasonable power engineering assumptions are incorporated. The output is available in tabular form or as graphical phase diagrams. Power Line Carrier: (Power) A high frequency (40-490 kHz), low power (1-10 watts) signal transmitted over power line conductors for system protection purposes and sometimes also used for communications and data links. Power Quality: (Power) Any power problem manifested in voltage, current, or frequency deviations that result in failure or misoperation of customer equipment.

18-68

Glossary

Precondition: To store information that will be acted on by an anticipated movement to cause a device or devices to function in a predetermined manner. Prime Ground Terminal: A terminal block or connection point in equipment housing to which all ground wires from equipment are connected and which acts as the interface point with wires to the main ground system. Probability Sample: A set of elements (sample) selected from a population under special requirements. Each element of the population must have a known non-zero probability of selection. Proximity Factor: A factor in the expression describing a complex shielding factor for two nested (or multiple) shields is called the proximity factor. This factor recognizes the interaction between the two shields and always makes the magnitude of the shielding factor larger (poorer shielding) than would be predicted on the basis of the product of the individual shielding factors. (See also Multiple Shields). Quadrupole: A system of two equal and oppositely directed dipoles constitutes a quadrupole. If the pair of dipoles are two-dimensional (parallel wires), the result is a two-dimensional quadrupole, if they are three-dimensional (current loops, the result is a three-dimensional quadrupole.) Similarly, there are even higher order multipoles. Quasistatic: A term used to denote the treatment of a time-varying field as essentially a static field because the frequency has a wavelength much larger than any dimensions under consideration. This is the case of the fields associated with 50/60 Hz electric power (and its harmonics) because these extremely-low-frequency fields (ELF) have a wavelength that is far larger than the maximum dimension of any practical object (the wavelength at 60 Hz is 5,000 km, or about 3,100 miles; for 50 Hz it is 6,000 km or about 3,700 miles). The main characteristic of Maxwells equations is that the electric field depends on the magnetic field through the time variation of the magnetic field, and that in turn the magnetic field depends on the electric field through the time variation of the electric field. Thus, the fields are mutually coupled, and this is the reason why, in general, we speak of an electromagnetic field, which propagates in space, as Maxwell showed, at the speed of light. This particular form of coupling disappears when the fields are time-invariant (static); the two fields can then be determined, independently of each other, from the free-charge distribution, the free-current distribution, and the properties of the medium. This is why static-field problems are relatively easy to solve. When the dimensions of the system under consideration are much smaller than the wavelength corresponding to the frequency of operation, a quasistatic approximation has sufficient accuracy for most practical purposes. For quasistatic fields (which oscillate relatively slowly), the coupling between fields is disregarded and both the electric field and the magnetic field are determined as if they were static fields. The reason that quasistatic electric fields can be treated as static fields is that in slowly varying electric fields, the time derivative of the electric flux density (the displacement current density) attains only minor (and insignificant) magnitudes, and the time rate-of-change of the displacement currents magnetic field is so small that the electric field induced by it is insignificant. A similar explanation can be presented for the magnetic field to be treated as a static field. Therefore, the electric and magnetic fields effectively do not influence each other at these low frequencies and can be de-coupled. Fortunately, for slow-varying fields, as any ELF field is, the problem is greatly simplified by resorting to approximate solutions of Maxwells 18-69

Glossary

equations, which lead to the treatment of quasistatic fields as static fields (i.e. electrostatic and magnetostatic) for most practical purposes. (See also Maxwells Equations). Radio Frequency Interference (RFI): RFI is intentional or unintentional electromagnetic energy that results in unintentional and undesirable responses from or performance degradation or malfunction of electronic equipment. Radio Noise: (Power) Electrical noise produced by corona on overhead transmission line conductors covering a large frequency spectrum that includes AM broadcast range from 500 kHz to 1.6 MHz. Radio noise from transmission lines is a design factor of transmission lines with voltages of 230 kV and above. It is present in both fair and foul weather, but is much greater in foul than in fair weather. The value of radio noise is usually reported as field strength, V/m, or in decibels referenced to 1V/m. If the transmission line geometry is modified to reduce electric and magnetic fields, radio noise may increase unless the conductor design is modified. Radio noise can also be produced by small gap discharges between loose or poorly fitting hardware. (See also Corona). Rail Joint Bond: (Historical) A metallic connection attached to adjoining rails to insure electrical conductivity. Railroad Grade Crossing: An intersection of two or more railroad tracks at the same elevation. Rated Voltage: The maximum voltage (rms) to which electric power system components may be subjected under normal operating conditions. Reactor: An electromagnetic device, the primary purpose of which is to introduce inductive reactance into a circuit. Receiver: (Railroad) (Train Control, Cab Signal, etc.) A device on a locomotive, so placed that it is in position to be influenced inductively or actuated by an automatic train stop, train control, or cab signal roadway element. Reclose/Reclosure: (power) The act of re-energizing a power line after protective relaying devices have tripped and de-energized the line. This may be done automatically or manually. Rectifier: (See battery charger or diode). Reduction Factor: (Field Reduction Factor) A term sometimes used to qualify shielding effectiveness. It is equal to the reciprocal of the shielding factor. Reflection: As an electromagnetic wave enters a conducting material, eddy currents are induced in the material. These eddy currents are essentially loops of currents (or dipoles) that create secondary electromagnetic waves that diffuse (inward and outward) away from the dipole source. The resulting secondary wave superimposes itself on the still-occurring primary electromagnetic wave and partial attenuation (due to reflection) occurs. Total attenuation happens only at superconductivity (in which case there is no absorption, but instead complete reflection of the

18-70

Glossary

wave). Reflection, for ELF magnetic field attenuation, is enhanced with increased conductivity. (See also Eddy Currents and Reflection Loss). Reflection Loss: The phenomenon of shielding reflection losses is analogous to reflection losses on a coaxial cable transmission line when it is terminated in an impedance other than its characteristic impedance. Reflection losses are better understood by considering electric fields separately from magnetic fields for power frequencies. It is simple to obtain high-performance shielding against electric fields at low frequencies because reflection losses are inherently high while absorption losses are low. Because most of the incident energy (due to the electric field) is reflected and little energy is transferred to the shield, absorption can be neglected. For magnetic fields, reflection losses are larger in materials with higher conductivity. When reflection losses are low, thicker highpermeability material is employed to increase magnetic shielding efficiency. (See also Reflection). Relative Permeability: In engineering applications the dimensionless ratio of the permeability ( ) of a material or medium with respect to the permeability of free space ( ) is sometimes used to quantify the materials magnetic properties. This ratio is call the relative permeability ( r ) since it is referenced to (or relative) to free space as follows:
r = , where = 4 x 10-7 weber/amp-m in the MKS (SI) system and 4 x 10-7 henry/m in

CGS:. (See also Permeability). Relay: (Railroad). Biased: (Biased Neutral) A relay that will operate to its energized position by current of one polarity only, and will return to its de-energized position when current is removed. Centrifugal: An alternating current frequency selective relay in which the contacts are operated by a fly ball governor or centrifuge driven by an induction motor. Code Following: A relay that will follow or reproduce a code without distortion within practical limits. Differential: A relay having windings operating in opposition. Double-Winding: A relay having two separate windings. Flasher: A relay so designed that, when energized, its contacts open and close at predetermined intervals. Frequency: A relay designed to respond to alternating current of a predetermined frequency. Light Out (LOR): A relay used to detect an open signal bulb filament. Line: A relay receiving its operating energy through conductors of which the track rails form no part. 18-71

Glossary

Magnetic Stick: A relay, the armature of which remains at full stroke in its last energized position when its control circuit is opened. Motor Type: A relay that operates on the principle of a motor. Neutral: A relay that operates in response to a predetermined change of the current in the controlling circuit, irrespective of the direction of the current. Overload: A relay that operates to open contacts when the current through its control coils exceeds a predetermined value. Polar: A relay which operates in response to a change in the direction of current in its controlling circuit and the armature of which may or may not remain at full stroke when its control circuit is interrupted. Polarized: A neutral relay equipped with polar armatures and contacts. Polyphase: An alternating current relay having two or more windings, operating on an induction motor principle, all windings of which must be properly energized. Power: A relay that functions at a predetermined value of the power. Power Transfer: A relay so connected to the normal source of power supply that the failure of such source of power supply causes the load to be transferred to another source of power supply. Quick Drop-Away: A relay which, when the controlling circuit is opened or completely shunted, will release quicker than an ordinary relay. Quick Pick-Up: A relay which, when energy is applied, will pick up quicker than an ordinary relay. Retained Neutral: A neutral relay, the armature of which is retained in the energized position for a predetermined interval of open circuit during the reversal of current in the control coils. Retained Neutral Polarized: A polarized relay, the neutral armature of which is retained in the energized position for a predetermined interval of open circuit during the reversal of current in the control coils. Single-Element: A relay, usually alternating current, having a single winding. Single-Winding: A relay having a single winding. Slow Drop-Away: (or Slow Release) A relay which, when the controlling circuit is opened or completely shunted, will release slower than an ordinary relay. 18-72

Glossary

Slow Pick-Up: A relay which, when energy is applied, will pick up slower than an ordinary relay. Thermal: A timing relay whose contacts are actuated by the heating effect of current flowing through its controlling element. Three-Position: A relay which operates in three positions. Timing: (or Time Element, of Timer) A relay which will not close its front contacts or open its back contacts, or both, until the expiration of a definite time interval after the relay has been energized. (See also Timer). Track: A relay receiving all or part of its operating energy through conductors of which the track rails are an essential part. Transformer: A relay in which the coils act as a transformer. Triple-Winding: A relay having three separate windings. Two-Element: A relay, usually alternating current, having two separate windings, both of which must be properly energized to cause the relay to operate. Two-Position: A relay which operates in two positions. Vane Type: A type of alternating current relay in which a light metal disc or vane moves in response to a change of the current in the controlling circuit. Vital: A relay constructed to have no wrong-side failures. Voltage: A relay that functions at a predetermined value of the voltage. Relay Cut Section: (See Cut-Section). Relay Type Interlocking: (Railroad) (Historical) An arrangement of signals, with or without other signal appliances, operated either from a control machine or automatically, and interconnected by means of electric circuits employing relays so that their movements must succeed each other in proper sequence, train movements over all routes being governed by signal indication. Release Value: The electrical value at which the movable member of an electromagnetic device will move to its de-energized position. Reluctance: A magnetic circuit can be compared with an electrical circuit so that the magnetic flux () corresponds to electric current (I), magnetomotive force (mmf) corresponds to electromotive force (emf), or voltage (V), and the opposition to the production of magnetic flux in a material (called its reluctance) is comparable with resistance (R). The symbol for reluctance is . Reluctance is inversely proportional to permeability (). As an example, iron has high 18-73

Glossary

permeability and low reluctance, whereas air or a vacuum has low permeability and high reluctance. The familiar Ohms law (V=IR) can be expressed for magnetic circuits: mmf (amp-turns) = . The mmf is considered to produce flux () in a magnetic material against the opposition of its reluctance (). Reluctance is defined as the flux path length (in cm) divided by the product of the materials permeability () and the cross-sectional area (in cm2). (See also Ampere-Turn and Permeability). Remanence: When a magnetizing field is applied to a specimen and then removed, some residual magnetization may remain (called remanence). In a typical magnetization curve or hysteresis loop, if the magnetic field, H, applied to a specimen is increased to a certain level and is then decreased, the flux density, B, decreases, but not as rapidly as it increased along the initial-magnetization curve. Thus, when H reaches zero, there is a residual flux density, or remanence, Br. (See also Coercive Force, Hysteresis and Retentivity). Repeater: A device conveying information as to the condition of an operated unit. Rephase: (Power) A term used to describe changes to power line phase conductor positions on a supporting structure(s). It is usually done to enhance field cancellation between adjacent circuits of a power line. An example of rephasing would be to change the phase positions on a double circuit transmission line to create opposite or cross-phasing for field management purposes. Rephasing is often done by changing the phase attachment points on the first structure outside substations at each terminus of a line. Sometimes rephasing of an entire power line route is done to achieve a field reduction at one location. Before rephasing a line, consideration is usually given to the initial cost to rephase and to potential impacts on maintenance, emergency repairs, any existing transpositions along the line route, changes in corona performance, different operation of the line in the future such that rephasing advantages disappear, and proximity to other power lines at different locations along a route at which rephasing could potentially increase field levels. Reset Device: (Train Control) A device whereby the brakes may be released after an automatic train control brake application. Resicalc: A residential magnetic field modeling program included as a module in the EPRI EMF Workstation. RESICALC is a Microsoft Windows application that allows users to map power-frequency magnetic fields in residential or neighborhood environments from user-defined arrays of transmission lines, distribution primaries and secondaries, and residential grounding systems. The most current version of RESICALC is 2.1. The program features an easy-to-use graphical user interface that is used to define complex residential settings with multiple power line and ground current sources. The residential model is used to calculate magnetic fields. The resulting magnetic field environment can be presented in a variety of graphical formats, including equi-field contour maps and three-dimensional surface maps of field intensity. RESICALC allows the user to perform sensitivity studies, produce multiple path linear profile plots, and to simulate taking spot measurements by using a field meter icon to display the magnetic field at specific locations in the model. Resonance: (Power) A situation where the capacitance and inductance of a circuit are equal and resistance is the only remaining impedance. 18-74

Glossary

Restoring Feature: (Railroad) An arrangement on a power-operated switch movement by means of which power is applied to restore the switch movement to full normal or to full reverse position, before the driving bar creeps sufficiently to unlock the switch with control lever in normal or reverse position. Resultant: (Power) A resultant is the square root of the sum of the squared rms values of the three orthogonal components that define an electric or magnetic field. The computed resultant can be larger than the actual maximum value the field attains. If the field is linearly polarized, the resultant and maximum field values will be the same. However, if the field is elliptically polarized the resultant will be larger than the maximum field because the resultant algebraically combines field components that do not achieve their maximum individual values at the same time. The largest error occurs for a field with circular polarization, in which case the resultant is about 41% larger than the actual maximum value the field can attain at any given instant. Many instruments for measurement of magnetic fields report a resultant value obtained from combining the rms value of the three individual components as described above. Retentivity: When a magnetizing field is applied to a specimen and then removed, some residual magnetization may remain (called remanence or retentivity). In a magnetization curve or hysteresis loop, the residual flux density, Br, when the magnetic field, H, reaches zero on the saturation loop is called retentivity (the materials ability to retain magnetization) or remanence (sometimes used to mean the ratio of the residual flux density Br, to the maximum flux density, Bm). (See also Hysteresis and Remanence). Retarder: A braking device built into a railway track to reduce the speed of cars. This can be done by means of brake shoes which, when set in position, press against the sides of the lower portions of the wheels. Group: A retarder that is so located that it is the last retarder the cars pass through before going into a single group of classification tracks. Inert: A non-powered non-releasable retardation device. Intermediate: A retarder that is located between master or hump retarder and group retarder. Master: A retarder or retarders located between the apex of the hump and the master switch or switches in a classification yard, and used specifically for car speed control. Pin Puller: A small retarder on the classification track side of the apex of a hump used to gather slack between cars to facilitate the uncoupling operation. Skate: An inert or weight responsive retarder at the pullout end of the classification tracks to prevent humped cars from fouling the pullout track. Tangent Point: A retarder located beyond tangent point at the entering end of classification tracks. Weight Responsive: A retarder that applies braking pressure proportional to the weight of the car. 18-75

Glossary

Retarder Control System: A system designed to control car movements, comprising car retarders together with such yard switch machines, skate machines, hump signals, trimmer signals and necessary control facilities as may be required. Reverse Phasing: (See Low Reactance Phasing). Reverse Position: (See also: Normal Position) The opposite to normal position. Rheostatic Braking: See dynamic braking. Right-Hand Rule: The Right-Hand Rule is used to determine the direction of a magnetic field around a conductor carrying a current. The conductor is grasped in the right hand with the thumb extending along the conductor pointing in the direction of the current flow. With fingers partly closed, the fingertips will point in the direction of the magnetic field. Right-Side Failure: (Railroad) The term right-side failure means that some significant deviation from normal signal system operation has occurred, but there has been no reduction in safety. This mode of operation may be due to a physical component failure in the signaling equipment, or a software defect, or some other cause. Regardless of the root cause, right-side failure modes are by far the most common cause of abnormal signal equipment operation. Most commonly, right side failures result from external track-related problems such as low ballast resistance, a broken wire, a broken rail, or some other readily identified physical failure. When modern signaling equipment enters a right-side failure mode, this means that the system has either detected an actual failure, or is otherwise unable to confirm its own proper functioning. The occurrence of a right-side failure means that the system has both detected the presence of a problem, and has taken appropriate action to ensure the continued safety of both train movements and the public at large. This action usually takes the form of placing the system in its most restrictive state, i.e. the wayside signals for the trains turn red, or the warning devices (gates, lights, bells, etc.) at a grade crossing are put into operation. The signal equipment will remain in this condition until the failure goes away, or possibly longer, due to the use of delay timers which prevent rapid back-and-forth transitions between the normal and failure modes of operation in the case of intermittent failures. In order to be vital, railroad signal equipment must enter a right-side failure mode whenever the proper functioning of any vital part of the signal system cannot be assured. This means that the equipment is intentionally designed to produce red wayside signals and false crossing warning system activations in response to the presence of poor ballast resistance, the presence of a track or wiring defect, or to the presence of interfering electrical signals on the rails. As inconvenient as this behavior may be, there is presently no safe alternative. (see also: Wrong-Side Failure) Riser: (Power) An electric power conductor that facilitates the transition between overhead and underground power lines. RMS: The root-mean-square (rms or RMS) value of a periodic function is the square root of the squared average value of the function taken throughout one period. Also called the effective value, the rms value of an alternating current is the number of amperes that, in a given resistance, produces heat at the same average rate as that number of amperes of direct current. Most electrical quantities

18-76

Glossary

used by electric power engineers and referred to in this handbook are usually given as the rms value (e.g. voltage, current, field strength), unless otherwise stated. Roundel: (See also Lens) A glass or similar product used in lens or reflector assemblies for spreading or deflecting, and/or coloring, the projected light beam into a pattern, dependent on the design. Route Locking: (See Electric Locking) The system in railway signaling whereby a route which has been set up for an approaching train is electrically and mechanically locked as the train approaches and while it passes through the route. The route is secured by the control system that prevents any conflicting routes being set up or signals being cleared. It is additionally secured by the passage of the train through the track circuits of the route concerned. Route Selection: (Automatic Switching for Classification Yards) This term is applied to a desired track destination established for an individual cut of cars by operation of a push button or other selective device. Route Signaling: The system common in the UK and throughout Europe of using specific aspects to designate certain routes as opposed to the practice in North America of using aspects to indicate speed over which a route must be traversed rather than the route itself. RSA: (Historical) Railway Signal Association. RS&I: (United States) Rules Standards and Instructions of the Federal Railroad Administration. RTC: (Canada) Railway Transport Committee. Running Rails: The two rails of a railway track upon which the wheels of the train rest and which provide the guidance for the train. SAE: Society of Automotive Engineers. Safe: (Railroad) The condition of a system, device or appliance that results in an indication or function equal to or more restrictive than is intended. Sag: The distance, measured vertically, from a span of a power line conductor to the straight line joining its two points of support. Saturation: The maximum value of magnetic flux density (Bm) in a material at a stated high value of field strength where a further increase in intrinsic magnetization with increasing field strength is negligible. This effect of little change in flux density (B) when field strength (H) increases is called saturation. Seam Sniffer: In the procedure used to test shielding effectiveness of an enclosure, the small-loop pickup device used to explore the seams of the shielding enclosure to test for magnetic continuity and search for defective seams is called a seam sniffer.

18-77

Glossary

Service Drop: (Power) The overhead service conductors from the last pole or other aerial support of the electric supply system connecting to the service entrance conductors at the building or service location. Service Entrance: (Power) The location where the neutral wire of the electrical service is connected to the grounding wire going to the grounding electrode. This location usually is at the electrical panel of a residence or at the main electrical panel of a building. Service Point: (Power) The point of connection between the facilities of the serving electric company and the premises wiring. Section Isolators: (UK) These are used on an electrified railway to divide the current supply system into separate areas or sections. Electrified railways are usually supplied from feeder stations (ac traction) or substations (dc traction) located at intervals along the line. Section Isolators are placed in trackside rooms at the boundaries between the feeds from two adjacent feeder stations. They are often referred to as Track Section Cabins. In addition to the track sectioning cabins, catenary isolators will be provided at strategic locations (crossovers and junctions) to facilitate maintenance and to minimize the operational impact of any catenary incident. These isolators will be either manually operated or motorized depending upon their respective safety and operational importance. Section Locking: (See Electric Locking). Sectional Release: A type of route locking in which directional stick relays unlock the route in sections. The purpose is to release switches or other devices in the route after the rear of a train movement has cleared them. Sectionalizing Switch: A switch for disconnecting a section of an electrical circuit from the source of energy. Selector Coil: (Railroad) (Historical) A coil which, when energized, will attract and hold in place an armature which in turn will permit a predetermined train movement to be made. Semaphore Arm: (Historical) The part of a semaphore signal displaying an aspect. It consists of a blade fastened to a spectacle. (See also Semaphore Blade). Semaphore Arm Spectacle: (Historical) That part of a semaphore arm which holds the roundels and to which the blade is fastened. (See also Spectacle). Semaphore Bearing: (Historical) A device that supports the pivot of a semaphore arm. Semaphore Blade: (Historical) The extended part of a semaphore arm which shows the position of the arm. (See also Semaphore Arm, Semaphore Arm Spectacle). Semaphore Counterweight: (Historical) A weight so connected that in case of breaking of the pipe controlling the signal the weight will fall and pull the signal to its most restrictive position. Semaphore Signal: (Historical) A signal in which the day indications are given by the position of a semaphore arm. 18-78

Glossary

Semi-Automatic Control: A control that is operated both manually and automatically. (See also Dual Control). Semi Automatic Signaling: A signaling system using track circuits as in automatic signaling, but with the ability of manual intervention to control trains or routes. With this method, large areas can be controlled semi-automatically using computers to route and regulate trains as well as record movements and log manual actions. The latest techniques involve GPS satellite location of trains. Service Box (Service Equipment): The necessary equipment, usually consisting of a circuit breaker or switch or fuses, and their accessories, located near the point of entrance of supply conductors to a building or other structure, or an otherwise defined area. It is intended to constitute the main control and means of cutoff of the supply and is constructed so that it may be effectively locked- or sealed. Sheath: A tubular impervious metallic protective covering applied directly over a cable core. Shed Receptacle: (UK) A socket provided on an electric railway vehicle where a shore (external) supply lead can be inserted to provide power when the normal traction supply is not available. Usually used in sheds and workshops. Shield: A conducting housing or screen composed of metal strands, ribbon or sheet metal that encloses a wire, group of wires, or cable, so constructed that substantially every point on the surface of the underlying insulation or core wrap is at ground potential or at some predetermined potential with respect to ground. The purpose of a shield is to substantially reduce the coupling of electric and/or magnetic fields into or out of circuits or prevents the accidental contact of objects or persons with parts or components operating at hazardous voltage levels. Shielding: An action using an enclosure, screen, or energized conductor(s) to reduce the strength of an electric or magnetic field in a given region. The materials used for shielding enclosures or screens are always metals, but there is a difference in using good conductors with low resistance or using good magnetic materials with high permeability like soft iron or mumetal. Energized conductors can also be applied to shielding problems as either active or passive conductors. The effectiveness of a shielding scheme is usually expressed as the shielding factor. (See also Active Shielding, Passive Shielding, and Shielding Factor). Shielding Effectiveness: A generic term, without a rigorous mathematical definition, referring to the ability of a shield to reduce field level. Shielding Efficiency: The ability of a given material or structure(s) to act as a shield against incident fields. (See also Shielding Factor). Shielding Enclosure: A structure that shields its interior from an external (or exterior) electric or magnetic field, or conversely, shields the surrounding environment from an interior electric or magnetic field source. A high-performance shielding enclosure is generally capable of reducing both the electric and magnetic field strengths by one to seven orders of magnitude, depending upon frequency. An enclosure is normally constructed of metal with provisions for continuous electrical contact between adjoining panels, including doors.

18-79

Glossary

Shielding Factor: An expression used to quantify the degree of shielding over a region of space is the shielding factor:
SF =
B , where B = the rms value of the magnetic flux density with the shield, and B = the rms B

value of the magnetic flux density without the shield. The shielding factor is a function of position, and in general, can also be a function of frequency, field strength, temperature, etc. Since the shielding factor is the ratio of two magnetic field values, it is a dimensionless quantity. In some literature, the shielding factor, sometimes called the attenuation, is expressed in decibels (dB) of attenuation with respect to the reference value without the shield as follows:
SFdB=20log( B ). B

Another related quantity sometimes used to describe shielding effectiveness is Shielding Efficiency (SE), where SE = (1 - SF) x 100%. (See also Transparency). Shielding Frequency: Any discussion of shielding often refers to the following ranges of frequency whose descriptions are drawn from standards for measuring the effectiveness of shielding enclosures: Power-Frequency: Low-Frequency: 15 Hz to about 10 kHz 14 kHz to 10 MHz

Medium-Frequency: 300 MHz to 1 GHz High-Frequency: 1.7 GHz to 6 - 18 GHz

Note: The names given to these frequency ranges (low, medium, high) are applied to shielding problems and are not the standard IEEE frequency ranges for other applications. Shield Wires: Wires that are grounded and strung above overhead transmission line conductors to provide protection from lightning. These shielding wires are commonly called overhead shield wires, overhead ground wires, or static wires. Shore Supply: (Railroad) An external power supply provided in a workshop or yard for vehicles that are disconnected from their normal traction or auxiliary supply. Special sockets or receptacles are provided on many railway vehicles for connecting shore (external) supplies. Sometimes referred to as shed supplies. See also knife switch. Short Circuiting Device: (UK) A hand tool for preventing traction current from being switched on to a dc-electrified line. Applies only to third rail systems. Generally mounted on trains and stations for use in emergency, they are not supposed to be used for discharging current, only for preventing recharge. A shorting tool is also used to ground or earth an overhead supply line by connecting the contact wire to the track rail in the section where it has been isolated and deenergized. 18-80

Glossary

Shunt: (Railroad) (United States) A by-pass in an electrical circuit. Shunts can be wideband, narrowband, and hardwire (used as a test or trial replacement component or a piece of equipment used to shunt the track). Sometimes also used for Termination Shunts. (See also Switch Shunting Circuit, Termination Shunt, Train Shunt Resistance). Shunt: (Railroad) (UK) To marshal vehicles in a given order to form a train. Shunt Wire: A wire forming part of a shunt circuit. See also Shunt (United States). Shunting Sensitivity: (Railroad) (See also Train Shunt Resistance) Shunting sensitivity of a track circuit is: Non-Coded Track Circuit: The maximum resistance in ohms that will cause the relay contacts to open when this resistance is placed between the rails at the most adverse shunting location. Coded Track Circuit: The maximum resistance in ohms that will prevent the code responsive track relay from following the code when this resistance is placed between the rails at the most adverse shunting location. SI: International System (of units). Siding: An auxiliary track for meeting or passing trains. SIG (Railroad) The abbreviation for signal. Signal: (Historical) (See Disc Signal, Semaphore Signal, Slotted Mechanical Signal, Smashboard Signal, Wig-Wag Signal). Signal (Application): (See also Interlocking Signals). Absolute: A signal of an automatic block signal system that is capable of displaying Stop as opposed to Stop and Proceed. Approach: A fixed signal used in connection with one or more signals to govern the approach thereto. Block: A fixed signal at the entrance of a block to govern trains and engines entering and using that block. CAB: A signal located in engine control compartment or cab indicating a condition affecting the movement of train or engine and used in conjunction with interlocking signals and in conjunction with or in lieu of block signals. Distant: A signal of fixed location indicating a condition affecting the movement of a train or engine. 18-81

Glossary

Fixed: A signal of fixed location indicating a condition affecting the movement of a train or engine. Note - The definition of a Fixed Signal covers such signals as switch, train order, block, interlocking, speed signs, stop signs, yard limit signs, slow signs, or other means for indicating a condition affecting the movement of a train or engine. Gate: (See Highway Grade Crossing Gate). Highway Grade Crossing: An electrically operated signal used for the warning of highway traffic at railroad-highway grade crossings. Holding: A fixed signal at the entrance of a route or block to govern trains or engines entering and using that route or block. Home: A fixed signal at the entrance of a route or block to govern trains or engines entering and using that route or block. Hump: A signal located near the summit in a hump yard that gives indication concerning movement to the classification tracks and indicates to the engineman the desired direction and speed of movement of his train. Permissive: A signal on which the most restrictive aspect is Stop and Proceed, or Restricting. Train Order: A signal used to indicate to a train whether or not it will receive orders. Trimmer: (or trim) A signal located, near the summit in a hump yard, which gives indication concerning movements from the classification tracks toward the summit. (See also: Interlocking Signals). Signal (Method of Control) Automatic: A signal controlled automatically. Non-Automatic: A signal controlled manually. Non-Stick: Opposite of Stick. Semi-Automatic: A signal that is controlled both manually and automatically. Stick: (Semi-Automatic) A signal so controlled that after automatically displaying a Stop aspect it will not again clear until its control lever is restored to normal and then to reverse, or until the control operator has performed other prescribed actions to permit a following train to proceed.

18-82

Glossary

Signal (Types & Arrangements): A means of conveying information for railroad operations. (See also Switch Indicator, Switch Position Indicator). Bracket: An arrangement whereby the signals for movements in the same direction on each of two or more tracks are mounted side by side on the same ground mast, using a cross piece rather than a cantilever arm. (See also: Mast). Color Light: A fixed signal in which the indications are given by the color of a light only. Color Position Light: A fixed signal in which the indications are given by color and the position of two or more lights. Dwarf: A low home signal. Flashing Light: A highway grade crossing signal, the indication of which is given by two horizontal red lights flashing alternately at predetermined intervals, or a fixed signal in which the aspects are given by color and by the flashing of one or more of the signal lights. Four-Position: A light signal unit arranged to provide four aspects. Light: A fixed signal in which the indications are displayed by the color or position of a light or lights, or both. Position Light: A fixed signal in which the indications are given by the position of two or more lights. Pot: A small revolving fixed signal used as a substitute for a dwarf signal. Searchlight: A type of color light signal that uses a single lamp with a single lens or lens doublet to display up to three different aspects by placing a color cone or disc between the lamp and lens. The desired color is selected by energizing an electromagnetic mechanism. The aspect displayed is dependent upon the polarity of the applied power. De-energization of the mechanism will cause the signal to display its most restrictive aspect. Three-Position: A semaphore arm or a light signal unit arranged to provide three aspects. Two-Position: A semaphore arm or a light signal unit arranged to provide two aspects. Signal Counter: (Historical) A device for registering the number of signal operations. Signaled Territory: This means any part of a rail system that has railroad tracks equipped with signal equipment to safely direct the movement of trains. The determination of signaled vs. nonsignaled depends only on the presence of a signal system for the trains, and does not depend on the presence or absence of grade crossing warning systems, or other signal equipment. (See also Dark Territory). Signalman: (UK) Person employed to operate or supervise the control of signals. Traditionally housed in a signal box, more recently a control room, where the signaling levers or controls are located. Also identified as a Signaler. Not to be confused with the same term used in the US for a signal maintenance person. 18-83

Glossary

Signal Mast: (See also: Mast) An upright support from which signals are displayed. Signals: (Railroad) (See Interlocking Limits, Interlocking Signals, Opposing Signals) Signals provide a visual indication passed to a train driver to advise the speed, direction or route of the train. There are almost as many types of signals as there are railways but they fall into the following main categories: Hand signals - used mainly where there are no fixed signals or where the fixed signaling has failed. Generally, each railway has its own defined hand signals recognized by its operators. Semaphore signals - a fixed track side signal where the stop indication is displayed as a horizontally positioned arm and proceed as a 45 or vertical arm. Color light signal - a fixed track side signal showing light indications to drivers. Cab signals - where the indications are displayed in the drivers cab. Six-Phase Transmission Line: (See High-Phase Order Transmission Line). Skin Depth: The depth in a material at which the current density (induced by an electromagnetic wave) is attenuated to 1/e (or 36.8%) of its value at the surface of the material. The skin depth is given by:
=
2

, where: = 2f, = permeability, and = conductivity.

As an example summarized below, consider the skin depth of a plane electromagnetic wave incident normally on a good conductor, such as copper ( = 1.26 H/m and = 58x 106 mho/m).
Frequency 60 Hz 1 MHz 30 GHz Skin Depth: Copper 8.5 mm (about inch) 0.66 mm 0.00038 mm

Thus, a high-frequency field is attenuated as it penetrates a conductor in a shorter distance than a low-frequency field - this phenomenon is often called skin effect. This is analogous to the way in which a rapid temperature variation at the surface of a thermal conductor penetrates a shorter distance than a slow temperature variation. (See also Absorption). Skin Effect: The phenomenon whereby a high-frequency field is attenuated as it penetrates a conductor in a shorter distance than a low-frequency field. It results in the nonuniform distribution of magnetic flux density in the cross section of the material. (See also Absorption and Skin Depth). Sleepers: (UK) In North America known as ties, short for crossties. The railroad components made of wood, concrete or sometimes steel, which are used to secure the rails at the correct gauge. Cast steel chairs fixed to the sleepers hold the rails in place by means of clips or keys. 18-84

Glossary

SLM: Switch-and-lock movement. Slotted Mechanical Signal: (Historical) A mechanically operated signal with an electromagnetic device inserted in its operating connection to provide a means of controlling the signal electrically as well as mechanically. Smashboard Signal: (Historical) A signal so designed that the arm will be broken when passed in the Stop position. Solenoid: A coil of conducting wire wound in a close-packed helix with a length much greater than the diameter of the coil. (See also Ampere-Turn and Electromagnet). Source Characterization: Field measurements and collection of all data necessary to specify the spatial and temporal characteristics of a field source are called source characterization. The level of detail required in performing a source characterization will depend on the final use of the information. Space Potential: The space potential of a point in an electric field is a phasor representing the voltage difference between that point and ground. The unit of space potential is the volt. The space potential is perturbed by the introduction of a conductive object in the field. Unperturbed space potential, which exists if the object is removed, is often used when a space potential value is given. Spacer Cable: (Power) A type of electric supply line construction consisting of an assembly of one or more covered conductors, separated from each other by insulating spacers and supported by a messenger cable. Span: The shortest distance between contiguous transmission or distribution line support structures. Spark Discharge: Discharges between conductive bodies at different potentials that come into contact in a power frequency electric field. For instance spark discharges may occur when a person touches a vehicle near a high voltage transmission line. The discharge occurs in the air gap just before the contact occurs. Often a spark discharge consists of a series of discharges as the arc extinguishes when the alternating voltage between the two bodies passes through zero and is reestablished when the voltage increases. Spark Over: A disruptive discharge between electrodes of a measuring gap, voltage-control gap, or surge protective device. SPD: Surge Protective Device. Spectacle: (Historical) That part of a semaphore signal which holds the roundels and to which the blade is fastened. (See also Blade Grip). Spectrum Analyzer: An instrument generally used to display the power distribution of an incoming signal as a function of frequency. Spectrum analyzers are useful in analyzing the characteristics of electrical waveforms in general since, by repetitively sweeping through the frequency range of interest, they display all components of the signal.

18-85

Glossary

Speed (By Rule) Limited: A speed not exceeding * miles per hour. Medium: A speed not exceeding * miles per hour. Reduced: Proceed prepared to stop short of train or obstruction. Restricted: Proceed prepared to stop short of train, obstruction, or switch not properly lined looking out for broken rail, not exceeding * miles per hour. Slow: A speed not exceeding * miles per hour. Yard: A speed that will permit stopping within one-half the range of vision. * Note: Railroads may insert definitions where asterisk is shown, suitable speed in miles per hour not exceeding 20 mph for Restricted Speed and/or Slow Speed, 40 mph for Medium Speed and 60 mph for Limited Speed. Split Phase (Power Line) : A single circuit three-phase power line in which one or more phases are split into two or more conductors (subphases). The phase current would divide among the sub-phases and, if the position of the subphases were chosen correctly, the magnetic field at ground level would be significantly reduced. Split-phase lines could have different configurations. Spring Switch: A track switch equipped with a spring device which forces the points to their original position after being trailed through and holds them under spring compression. Spring Switch Protection: An arrangement of circuits whereby proceed aspects of a signal cannot be displayed unless the switch and its controlling lever or equivalent device are in corresponding position. Squeeze-Out: (Railroad) A term generally used in conjunction with retarders, having reference to a situation where car wheel or wheels are lifted off track and ride retarder brake shoes due to excessive pressure as related to weight of car. SSAC: Steel Supported Aluminum Conductor. Standard Code: (of operating rules) The operating, block signal and interlocking rules of the Association of American Railroads. Station: (Railroad). Block: A place designated by timetable at which manual block signals are displayed. Control: The place where the control machine of a traffic control system is located. Interlocking: A place from which an interlocking is operated.

18-86

Glossary

Steady-State: A condition in which voltages and currents in an electrical system do not vary appreciably with time. Voltages and currents may be alternating, in which case the steady-state condition refers to their rms values. Intermittents and transients are deviations from the steadystate conditions. The term steady-state applies to electric and magnetic fields as well. Step Potential: The potential difference between two points on the earths surface separated by a distance of one pace (assumed to be one meter) in the direction of maximum potential gradient. This potential difference could be dangerous when current flows through the earth or material upon which a worker is standing, particularly under fault conditions or lightning strikes to earth. Stepper: A group of relays and circuits in a code system that receives and interprets the control codes and generates indication codes from the field to the office. Stick Circuit: A term applied to a circuit used to maintain a relay or similar unit energized through its own contact. Also called a stick path. Stick Signal: (See Signal). Stinger: A North American term for a shore (external) supply lead used to provide power to electric trains inside workshops. Stop Indication Point: (Train Stop or Train Control) As applied to an automatic train stop or train control system without the use of roadway signals, a point where a signal displaying an aspect requiring a stop would be located. Stopping Distance: (See Braking Distance). Style C: (See Type C). Subcalc: A substation magnetic field modeling program included as a module in the EPRI EMF Workstation. SUBCALC is a Microsoft Windows application that allows users to estimate power-frequency magnetic fields in and around electric power substations based on models of the facilities. The most current version of SUBCALC is 2.0. The program features an easy-to-use graphical user interface that is used to define substation equipment and power lines used in creating a model for calculation of magnetic fields. The resulting magnetic field environment can be presented in a variety of graphical formats, including equi-field contour maps and threedimensional surface maps of field intensity. SUBCALC allows a user to carefully map the spatial characteristics of the magnetic field and to determine how line or equipment configuration, electrical characteristics, or location might affect field levels at locations of interest inside a substation or in the nearby region. Sub-Distribution Panel: (Power) An electrical panel in a building providing electrical power to several other panels and receiving its power from the main distribution panel. Super Bundle Phasing: A term that describes multiple electrical circuits arranged in a manner so that the phase sequence for one circuit is placed adjacent to the same phase sequence for the other circuit. For example, for a double circuit line with vertical configuration, one circuit may 18-87

Glossary

have A-B-C phases arranged top to bottom and the adjacent circuit would also have A-B-C arranged top to bottom. This method produces somewhat higher fields at ground level than low reactance or cross phasing, but a super bundle configuration has somewhat lower conductor surface gradients with less potential for corona related phenomena. Super bundle phasing is also called like phasing and similar phasing. For the case where current flow in adjacent circuits is in opposite directions, a super bundle will produce lower ground level magnetic fields than low reactance phasing. (See also Low Reactance Phasing). Surge: A transient voltage or current, which usually rises rapidly to a peak value and then falls more slowly to zero, occurring in electrical equipment or networks in service. Surge Arrester: (See Surge Protective Device). Surge Arrester Gas-Tube: A gap or series of gaps in an enclosed discharge medium designed to protect apparatus or personnel or both from high transient voltages. Surge Impedance: Surge impedance of a transmission line is the impedance of a load that would cause no net reactive power flow and approximately a flat voltage profile along the line length. For a single-phase lossless line, the surge impedance is equal to:
Zs = X L Yc , where XL and YC are the series inductance and the shunt admittance per unit of

length of the line, respectively. If line losses are neglected, the surge impedance is a pure resistance. The surge impedance loading of a line (SIL) is the power carried by the line terminated in its surge impedance as follows:
SIL = V2 ZS

, where, for a three-phase line, V is the line voltage. Surge impedance loading is an

important parameter affecting the maximum electrical load of a line in stable steady-state conditions. Surge Let-Through: The peak voltage level of the surge that passes by .a surge protective device. Surge Protection: (See Lightning Protection) Protection to equipment and personnel from high transient voltages such as those caused by lightning or other abnormal conditions. Surge Protective Device (SPD): The generic term used to describe a device by its protective function, regardless of technology, uses, ratings, packaging, point of application, etc. An SPD is intended to either limit transient over voltages or divert surge currents or both. It contains at least one nonlinear component. Also called arrester, lightning arrester, surge protector, and surge arrester. Surge Protective Device Mode(S) of Application: A mode of operation for an SPD identified as having a protective element or elements connected in the following applications: l) Across pairs of energized (non-grounded) lines identified as Line-Line. 2) Between a non-grounded conductor and a neutral conductor, identified as Line-Neutral. 3) Between a neutral conductor and grounded conductor, identified as Neutral-Ground. 4) Between a signal conductor and ground, identified as Line-Ground. 18-88

Glossary

Note: A surge let-through voltage level, maximum continuous operating voltage level, and maximum surge current level shall be identified by each mode of operation of the SPD. Follow through current performance levels shall also be identified for all operational modes of arrestor gas and air gap type SPDs. One may also want to identify maximum (electric power company) fault current levels for each mode of operation of a series operated SPD. Surge Protective Device, Primary: A surge protective device applied on a circuit at its entrance to a facility and rated to withstand the severest category of impressed surges for that environment. Surge Protective Device, Secondary: A surge protective device applied on a circuit usually protected by a primary surge protective device and rated to withstand the severest category of impressed surges for that environment. Typically located away from the circuits entrance to the facility. (Note: In areas of low exposure, a secondary surge protective device might be used in lieu of a primary surge protective device). Surge Protective Device, Tertiary: Components integrated into the design of the device they are intended to protect. Operating parameters of tertiary protection need to compliment the characteristics of any upstream surge protective devices (i.e. primary and secondary). Surge Protective Device, Signal: A surge protective device specifically designed for circuits used in railroad signal systems. Surge Protective Device, Vital: A primary surge protective device intended for use in railroad signal circuits that cannot be grounded without reducing safety. Surge Protector: See Surge Protective Device (SPD). Surge Withstand Capability: The maximum dc, ac rms, or impulse voltage and energy which can be applied between terminals of a circuit, or between a circuit and metallic chassis of the equipment, without resulting in functional, electrical or mechanical damage to the equipment. Susceptibility-Electrical Circuit: (Power) The characteristics of a circuit and its peripheral devices that determine the extent to which it is adversely affected by inductive fields. Susceptibility- Magnetic Property: Many materials respond to an applied magnetic field by becoming magnetized to some degree. The degree to which they become magnetized depends on the strength of the applied field and on the characteristics of the material. The ratio of the intensity of magnetization (M) to the applied magnetic field (H) is a dimensionless property of a material called susceptibility. (Note: Both H and M have the same units of A/m). Susceptibility represents the ability of a material to be magnetized. If the susceptibility , is a constant (not a function of the applied field, H), the material is said to be a linear material such that: M=H. For most paramagnetic and diamagnetic materials, susceptibility is very nearly constant over a wide range of magnetic field levels and is therefore considered to be linear. Paramagnetic materials have a relatively small positive value for susceptibility and diamagnetic materials have a relatively small negative value. Ferromagnetic materials are generally non-linear. (See also Magnetization and Permeability). 18-89

Glossary

Switch: (Railroad) A pair of switch points with their fastenings and operating rods providing the means for establishing a route from one track to another. (See also Facing Point Switch, Interlocked Switch, Trailing Point Switch, Turnout). Double-Slip: A combination of a crossing and two connecting tracks, located within the limits of the crossing, each being made up of a right-hand switch from one track and a left-hand switch from the other track, which unite to form the respective connecting tracks without additional frogs. Dual Control: A power-operated switch also equipped for hand operation. Power-Operated: (See also Dual Control) A switch operated by some form of energy, usually electrical or pneumatic. Single-Slip: (See also Dual Control Switch) A combination of a crossing and a single connecting track, located within the limits of the crossing, and made up of a right-hand switch from the one track and a left-hand switch from the other track, which unite to form the connecting track without additional frogs. Spring: A switch equipped with a spring mechanism arranged to restore the switch points to normal position after having been trailed through. Switch-And-Lock Movement: A device, the complete operation of which performs the three functions of unlocking, operating, and locking a switch, movable point frog, or derail. (See also Switch Machine). Switch Heater: A winter switch protection device consisting of either gas combustion burners or electric heating elements fastened directly to the stock rail or switch point in which the heaters raise the temperature of the steel to melt snow and ice. Gas combustion burners may be of the direct-flame impingement or radiant designs. Electric heating elements may be of the tubular (Calrod), plate, or pad designs. Switch Indicator: An indicator used at a non-interlocked switch to indicate the condition of a block. (See also Block Indicator, Switch Position Indicator). Switch Lever Lock: (Railroad) (Historical) An electric lock connected to a lever in an interlocking machine to prevent the movement of lever or latch until released. Switch Machine: (See SLM, Yard Switch Machine). Switch Point: A movable tapered track rail, the point of which is designed to fit against the stock rail. Switch Position Indicator: A low two-aspect horizontal color light signal with electric lamps for indicating position of switch or derail.

18-90

Glossary

Switch Shunting Circuit: (Railroad) A shunting circuit that is closed through contacts of a switch circuit controller. Switch Target: A device mechanically actuated by a switch stand, or a switch point, to indicate the position of the switch. Symmetrical Components: Symmetrical components is a mathematical technique developed by Mr. C.L. Fortescue (published in 1918) that can be applied to a variety of engineering problems. It is widely used for solving power engineering problems involving unsymmetrical (or unbalanced) power systems. In three-phase power systems it is applied to current, voltage, and impedance problems. In this reference book, the method is applied usually to circuit or load condition issues that could affect EMF levels and attenuation characteristics. In general, an unbalanced system can be resolved into balanced, symmetrical three-phase or single-phase vector systems (called symmetrical components) and used to obtain solutions to the original unbalanced, non-symmetrical problem. The unbalanced system is uniquely resolved into three sets of balanced phase-sequence vectors called: positive sequence, negative sequence, and zero sequence components. In the method of symmetrical components all of the resolved phase-sequence vectors rotate in the positive (counterclockwise) direction. The three vector components of the positive sequence components are equal to each other in magnitude, are 120 electrical degrees apart in phase, and achieve maximum values in the positive phase rotation sequence of a, b, c. The three negative sequence vector components are also equal to each other in magnitude, are 120 electrical degrees apart and rotate counterclockwise, but in the negative phase sequence of a, c, b. The zero sequence vector components are likewise equal to each other in magnitude, but all three components are in phase (i.e. there is a zero degree angle or zero sequence between the three components). The magnitudes of the components within each of the three phase sequences (positive, negative, zero) are equal but they may or may not be equal to each other in magnitude. And the rotating vectors of each of the three sets of components may be shifted by some electrical angle from a common reference point in the rotating vector diagrams. A physical interpretation of positive sequence components, for example, would be the currents that would occur on a balanced power system; the sum of their instantaneous values is zero. Negative sequence components would exist (along with positive sequence components) in an unbalanced system in which the phase current magnitudes are numerically unequal but still sum to zero. An example could be unequal phase currents due to differences in phase capacitance in a long horizontal configuration transmission line (i.e. the center phase would have larger charging currents). Zero sequence currents exist when there is a net current (i.e. the instantaneous values of the phase currents do not sum to zero). The zero sequence phase currents would flow in the neutral or ground paths. The EPRI Power Line Calculator program uses the method of symmetrical components to solve unbalanced conditions. Symmetric Voltages: Three-phase line voltages are symmetric when the phase angle between them is 2/3 radians (120 degrees). Transmission line voltages are generally considered perfectly symmetric. Synchronous Motor: A synchronous machine that transforms electric power into mechanical power. Unless otherwise stated, it is generally understood that a synchronous generator (or motor) has field magnetics excited with direct current.

18-91

Glossary

Take Siding Indicator: An indicator generally used to convey instruction to approaching trains to take siding. (See also Leave Siding Indicator). Talker: A device generally used in conjunction with defect detectors to communicate a verbal message by radio to trains. Tamping: The process by which ballast is packed around the sleepers of a track to ensure the correct position for the location, speed and curvature. Can be done manually or mechanically by special tamping machines, usually independently powered track vehicles. Tangent Point: The point of transition between straight and curved track. Tem: A mode whose electric and magnetic field vectors are both perpendicular to the direction of propagation is called the Transverse ElectroMagnetic mode (TEM). No component is in the axial direction. Termination Shunt: (Railroad) An electrical device used to define the end of an approach to a grade crossing, which works by providing a low-impedance electrical path from one rail to the other on a railroad track. This may consist of a length of wire, a very large (>100,000 F) nonpolarized capacitor, or a tuned (series-resonant) inductor-capacitor (LC) combination. Tesla: The tesla is a unit in the MKS (SI) system used to describe magnetic flux density (B), or magnetic flux lines per unit of cross-sectional area. The tesla is 1 weber per square meter (or 8 2 4 1x10 lines/m ). The tesla is equal to 10 gauss in the CGS system: 0.1T = 1mG. (See also Flux, Magnetic Flux Density and Weber). Third Rail System: Traction current collection system that uses an additional rail to transmit the electrical supply to the train. See Conductor Rail. Three-Way Switch: (Power) A three-way switch is a wiring arrangement such that an electrical load can be operated from two or more different switches (for example, two light switches at opposite ends of a long hallway to control a ceiling light in the hallway center). In a three-way switch configuration, there is usually a third wire between the switches (in addition to the single hot leg and the neutral wire). This third wire is usually another hot leg (sometimes called a carrier leg). If the neutral wire or either of the hot legs is located along a path separate from each other, then a loop current flow can create an elevated localized magnetic field whenever the electrical load is turned on. Thyristor: A solid-state power electronic device for the control or switching of ac power where switching, multicycle control, and phase control are included. The only power-controlling element is the thyristor, although other power elements may be included. The term siliconcontrolled-rectifier (SCR) was previously used to describe this device, but in 1963, IEEE officially adopted thyristor as its name for the SCR. The name thyristor is a combination of thyraton and transistor to create thyristor. The thyraton was a 1921 modification of the mercuryarc valve (vacuum tube) by GE. Thyratron was formed from the Greek words thyra (gate) and tron (tube).

18-92

Glossary

Time Release: (Railroad) (See also Locking) A device used to prevent the operation of an operative unit until after the expiration of a predetermined time interval after the device has been actuated. To Clear A Signal: (Railroad) To permit or cause a signal to display an aspect, the indication of which is more favorable than STOP. Total Harmonic Distortion (THD): For a complex waveform, the total harmonic distortion is the ratio of the rms value of the sum of the squared individual harmonic amplitudes to the rms value of the fundamental frequency. Touch Potential: The voltage potential difference between a conductive structure and a point on the earths surface separated by a distance equal to the normal maximum horizontal reach, approximately one meter. This voltage potential difference could be dangerous and could result from induction, lightning or fault conditions. Towers: (Power) Electric power line towers are usually made of galvanized or painted steel, aluminum, or treated wood members. Track Block: A block into which a signal cannot be cleared. Track Blocking: Track blocking is a method of preventing the clearing of any controlled signal governing movements into a block section of track. Track Circuit: An electrical circuit that uses the track rails as the conductors between a transmit and receive device, the limits of which are commonly defined by the location of insulated joints. The primary purpose of the track circuit is to detect an occupancy or interruption. It may also be used to convey information. AC: A track circuit using alternating current source, generally operating at or below 200 Hz. Audio Frequency: An electronic track circuit, generally modulated, operating in the audio frequency spectrum. This type of track circuit does not require use of insulated joints to define its limits. Clearance: A track circuit used to detect when cars have reached or cleared a predetermined point commonly used in classification yards. Coded: A track circuit in which the energy is pulsed at predetermined rates. DC: A track circuit that uses a low voltage direct current source. Detector: A track circuit used to detect track occupancy in a specific location. Distance To Couple: A track circuit used to measure the distance between the clearance point after the last switch and the nearest standing car on the classification track. 18-93

Glossary

Electronic: A circuit which performs the functions of a track circuit using electronic devices to transmit and receive information in the form of dc or ac pulses or at ac modulated frequencies. High Level AC DC: Generally referring to a track circuit that employs relatively high alternating current voltage on rails, low impedance energy source, and transformer-rectifier unit between rails and direct current track relay. Island: A short, defined track circuit with spans a highway grade crossing. Motion Sensitive: An audio frequency track circuit, generally operating below 1000 HZ, primarily used to activate highway grade crossing warning systems. Overlay: An electronic track circuit, generally superimposed over an existing track circuit. Phase Selective: A coded track circuit using a phase selective unit. Primary-Secondary: A special arrangement of dc track relay and a repeater relay used for increasing shunting sensitivity. Track Circuit Connector: A device used for connecting one or more wires to a rail. Track Gauge: The distance between the inner faces of the railheads of a railway track, commonly referred to as the gauge. Track Indicator: (See also Block Indicator, Switch Indicator) An indicator used to indicate the condition of a given track section. Track Indicator Chart: (Historical) A map-like reproduction of railway tracks controlled by track circuits so arranged as to indicate automatically, for defined sections of track, whether or not such sections are occupied. Track Instrument: (Historical) A device in which the vertical movement of the rail or the blow of the wheel operates a contact to open or close an electrical circuit. Track Shunt: (See Shunt). Track Switch: (See Switch). Track Wire: A wire used to connect signaling equipment to the rails. Traction Current: Term used for electric power supply used on electric railways for trains. Normally supplied by overhead wire or third rail. Traction Motor: Electric motor used to provide the final drive to a locomotive or train axle. Used in diesel-electric and electric systems. The traction motor is mounted close to the axle and transmits power through a reduction gearbox. 18-94

Glossary

Tractive Effort: The power that a locomotive is able to exert before the wheels slip out of control. It is calculated by multiplying the weight on the driving axles by the coefficient of adhesion. Traffic Control System: A block signal system under which train movements are authorized by block signals whose indications supersede the superiority of trains for both opposing and following movements on the same track. (See also Centralized Traffic Control). Traffic Lever: A lever, or equivalent controlling device used as a check lever, crossing lever, detector lever, master lever, route lever, also to control another lever, group of levers or functions to establish traffic direction. Traffic Locking: (See Electric Locking). Trailing Movement: The movement of a train over the points of a switch that face in the direction in which the train is moving. Trailing Point Switch: The points of a track switch that face away from traffic approaching in the direction for which the track is signaled. Trailing Release: (See Sectional Release). Train: An engine or more than one engine coupled, with or without cars, displaying markers. Train Control System: A system so arranged that its operation would automatically result in the following: A full service application of the brakes that will continue either until the train is brought to a stop, or, under control of the engineman, its speed is reduced to a predetermined rate. When operating under a speed restriction, an application of the brakes when the speed of the train exceeds the predetermined rate and which will continue until the speed is reduced to that rate. Train Shunt Resistance: The actual resistance (in ohms) from rail-to-rail through wheels and axles of a train, engine or car. This resistance will vary with rail and wheel surface conditions and with weight of equipment. Train Stop System: A system so arranged that its operation will automatically result in the application of the brakes until the train has been brought to a stop. Transformer: This device transfers electric power from one circuit to another by electromagnetic coupling through a ferromagnetic (iron) core, usually with changes of both voltage and current. A step-up transformer receives electrical power at one voltage and delivers it at a higher voltage. A step-down transformer receives electrical power at one voltage and delivers it at a lower voltage.

18-95

Glossary

Transformer Vaults: (Power) Electric power transformers that are installed inside a building or under sidewalks are usually placed in a reinforced concrete vault. Transient: (Power) A magnetic field transient occurs during a period of time (e.g. a small fraction of one cycle) when the field has rapid changes in values while passing from a steady state condition to another. The most important aspect of a transient is the rapidity of the change. A transient is such that the derivative of the magnetic field versus time (db/dt) reaches a peak significantly greater than the peak value of db/dt in the steady state conditions preceding and following the transients. Transient: A change in the steady-state condition of voltage or current, or both. Transients may be caused by lightning, a fault, or by power line switching operations, and may readily be transferred from one conductor to another by electrostatic or electromagnetic coupling. The time duration of a transient is often recognized as an event with duration of less than 8 milliseconds as differentiated from a surge which is an event with duration greater than 8 milliseconds. Transient Voltage Surge Suppressor (TVSS) (See Surge Protective Device, SPD). Transition Structure: (Power) A pole or structure that facilitates the transition between overhead and underground electric power lines. Transmission Coefficient: An expression of shielding effectiveness, called the transmission coefficient, is the ratio of magnetic flux density with shielding present to flux density without shielding. (See also Shielding Factor and Transparency). Transmission Line Reference Book: 345 Kv And Above: An invaluable power engineering reference book produced by EPRI (sometimes called the red book due to its cover). The second edition is the most recent version. The book is an excellent source of technology and data based largely on experimental work performed over about two decades at EPRIs high voltage research center in Lenox, MA. Four major design areas are covered: insulation design, corona performance, electric and magnetic fields, and circuit performance (including conductor characteristics). In each area, a common set of base-case circuit designs is analyzed to help the design engineer understand the relative importance of various parameters affecting design and performance. Transparency: The extent to which a material or medium allows magnetic flux to pass unaffected through it - the reciprocal of shielding factor. (See also Reluctance and Shielding Factor). The following table gives some practical examples of transparency:
Object Faraday Cage (iron mesh spaced at 3 cm) Volkswagen Iron Sheet Garage Sleeping Bag (with copper 0.1% layer, buttoned up) E inside/E outside 0.5% 1.0% 0.1% B inside/B outside 65% 50% 7% 90%

18-96

Glossary

Transposition: (Power) A term used to describe the interchange of transmission lines phase conductor (and possibly overhead shield wire) positions at selected intervals (often at the onethird points along a lines length or at a specified distance) to compensate for inductive effects. Transverse Mode Voltage: (See Differential Mode Voltage). Trap Circuit: A term applied to a circuit used at locations where it is desirable to protect a section of track but where it is impracticable to maintain a track circuit. Also an early form of constant warning equipment used at highway crossings. (See also Dead Section). Trunking: (UK) (Historical) A casing used to protect electrical conductors. Turnback Track: A North American term for a reversing track or center siding. Turnout (UK): (See also Points.) Track work where a single track splits to become two tracks and equipped with moving rails to change the route. Also referred to as points in the UK and a switch in North America. Type C: A track circuit that uses an alternating current source and a dc detection relay at the feed end and a diode between the rails at the far end. (Also referred to as a Style C or ac/dc). UG Underground. Referred to underground power lines. UHV: Ultra High Voltage. Transmission line voltages of 1000 kV or greater are considered UHV lines. No such voltage exists in North America. Japan and Russia have built sections of lines that may operate at about 1100 kV in the future. UIC: Union Internationale de Chemin de Fer - International Union of Railways - the French dominated European railway regulating body which sets engineering and operating standards for railways. Equivalent in function to AAR and AREMA in North America. Ultrasonic Flaw Detection: A system for examining the condition of rails to determine the integrity of the steel. An electronic instrument that is run along the track collects the data. Also used to detect flaws in railway axles. Unbalance: (Power) Unbalance of a three-phase distribution line is a measure of the deviation of the individual phase currents from the average phase current. Most commonly the unbalance, U, is defined as the maximum deviation between a phase current and the average phase current divided by the average phase current as follows:
U= I p I av I av

, where Ip is the phase current that deviates the most from the average current, Iav.

Unbalance has also been defined as a current, equal to the vectorial sum of the three phase currents. If the phase currents are balanced, the unbalance is zero. In a system with neutral, the neutral is designed to carry the unbalance. However, part of the unbalance often flows elsewhere and the distribution line exhibits a net current. 18-97

Glossary

Underbuild: (Power) A term used in transmission line engineering to describe lower voltage circuits placed below and on the same structure as the phase conductors of a transmission line. Uniform Field: A region has a uniform field if, in all points of the region, the magnitude and direction of the field are constant. Unlike Phasing: (See Low Reactance Phasing). Unperturbed Field: The field in the vicinity of an object may be perturbed by the presence of the object. The unperturbed field is the field that is present when the object is removed. Because the electric field at or close to the surface of an object is generally strongly perturbed, the value of the unperturbed electric field is often used to characterize the field and to evaluate electric field effects of transmission lines and stations. Both ferromagnetic and conductive objects can perturb AC magnetic fields. Only ferromagnetic objects perturb DC magnetic fields. Unsafe: (Railroad) The condition of a system, device or appliance that results in an indication or function less restrictive than is intended. URD: Underground Residential Distribution. Vertical Configuration: (Power) A configuration in which the three phases of each three-phase circuit of the line are stacked vertically one above the other (the conductors may not always lie in the same vertical plane due to offsets to reduce conductor motion problems caused by sudden ice-drop or wind-induced galloping). Vibrating Bell: (Historical) An audible signal which, when once started, continues automatically until the circuit is opened. Virgin Iron: This is iron that has not yet been strongly magnetized. The term is used in recognition that the magnetization characteristics of iron (and other ferromagnetic materials) can depend on its prior magnetization history. (See also Hysteresis and Magnetization Curve). Vital Circuit: (Railroad) Any circuit the function of which affects the safety of train operation. (See also Fail Safe). Vital Relay: (Railroad) A relay, meeting certain stringent specifications, so designed that the probability of its failing to return to the prescribed state upon de-energization is so low as to be considered practically nonexistent. VLF: Very-low-frequency, defined by IEEE as the frequency range from 3,000 to 30,000 Hz (330 kHz). Voltage- Circuit: (Power) The greatest rms difference of potential between any two conductors of the circuit (also called line-to-line voltage).

18-98

Glossary

Voltage Gradient: A vector equal to and in the direction of the maximum space rate of change of the voltage at a specified point. Voltage gradient is synonymous with potential gradient and is often referred to simply as gradient or field strength. It has units of volts per meter. Voltage- Nominal: (Power) A nominal valve assigned to a circuit for the purpose of conveniently designating its voltage class. The actual voltage at which a circuit operates can vary from the nominal within a range specified by the electric company and that permits satisfactory operation of electric power equipment and attached loads. Voltage Range: (Railroad) (Historical). See also Battery (Voltage). First: (Historical) Thirty volts or less. Second: (Historical) Over 30 volts to and including 175 volts. Third: (Historical) Over 175 volts to and including 250 volts. Fourth: (Historical) Over 250 volts to and including 660 volts. Fifth: (Historical) Over 660 volts. Voltage-To-Ground: (Power) The rms voltage between a specified conductor and the point or conductor of the circuit that is grounded (also called line-to-ground and line-to-neutral). VVVF: Variable Voltage Variable Frequency traction drive system. Used where 3-phase ac motors are provided on rolling stock. Wavelength: The length of one complete cycle of an electromagnetic wave (i.e. the ratio of the speed of light to frequency). The distance is measured in the direction of propagation between a reference point on one wave and the next point that has exactly the same relative phase. Wayside: The area of a railroad right-of-way adjacent to the tracks. Weber: The unit of magnetic flux is the weber in the MKS (SI) system. One weber (Wb) equals 8 1x10 magnetic lines of force. In the CGS system, the unit of magnetic flux is the maxwell (Mx), and one Mx equals one line. Since the weber is a large unit for typical fields, the microweber unit can be used (1Wb = 100 lines or 100 Mx). Westinghouse Brake: Widely used automatic air brake invented in the 1870s by George Westinghouse and developed worldwide. Wheel Detector: A device capable of detecting the presence or passage of a wheel. (See also Axle Counter). Wheelset: A fixed formation of an axle with two wheels set at the correct gauge for the track.

18-99

Glossary

Wheel Slide: Synonymous with skidding and usually caused by over braking during poor adhesive conditions. It is a common cause of wheel damage, as it produces a flat spot (called a flat) on the wheel where the skid occurred. Severe flats have been known to derail a train. Modern rolling stock is equipped with various systems to assist with the elimination of wheel slide. These include load control, automatic brake dumping if a slide is detected, rail applications like Sandite to improve adhesion and attention to maintenance of correct mechanical brake settings. Wheel Slip: Phenomenon caused on a locomotive or power vehicle by over-application of power to the drive system relative to the available adhesion. It can cause damage to electric motors and is normally automatically detected to immediately eliminate or reduce the power being applied. White Light: (Historical) A term referring to the color of the light from a light unit using a clear uncolored lens or roundel in combination with a kerosene or electric illuminant. (See also Bare Illuminant). Wig-Wag Signal: (Historical) A highway crossing signal, the indication of which is given by a horizontally swinging disc with or without a red light attached. (See also Banner). Winter Switch Protection Device: (Railroad) A device used at track switch and moveable point frog locations where it is necessary to prevent snow and ice from interfering with the movement of switch points. Withstand Voltage: The maximum value of voltage applied to electric power system insulation that does not result in a flashover or insulation failure. Working Value: (Railroad) The electrical value which, when applied to an electromagnetic instrument, will cause the moving member to move to its full-energized position to provide maximum front contact pressure. (See also Pick Up Value). Wrong-Side Failure: (Railroad) The term wrong-side failure means that a deviation from normal signal system operation has occurred, the signal system has not detected it, and this has resulted in a potentially unsafe condition or mode of operation. As the purpose of most railroad signaling equipment is to provide an indication of what movements may safely be made, both to trains and to the motoring/pedestrian public, the end result of this is usually some form of a false clear signal aspect, which gives the wrong indication to the train operator or public. The term false clear comes from the terminology used to describe wayside signal aspects. The proper name for a green signal aspect is clear. If a train has a clear signal, this means that they are looking at a signal that is usually green in color, and tells them that it is safe to proceed onto the track(s) ahead, at or below the speed limit. Abnormal operation resulting from wrongside failures is extremely rare. However, it is important to note that in the case of severe railroad signal equipment damage resulting from a catastrophic event, multiple vital component failures may be induced in the signaling equipment simultaneously. This most often results from lightning, surges, or power line faults that cause excessive voltage to be applied to the inputs or outputs of the signal equipment. (see also: Right-Side Failure).

18-100

Glossary

WYE Connection: A wye connection is a method of connecting the windings of a transformer so that one end of each of the windings is connected to a common neutral point and the other end of each winding is connected to its appropriate load. This type of connection is also called a Y or star connection. Yard: (Railroad) A system of tracks within defined limits provided for the making up of trains, storing of cars and other purposes, over which movements not authorized by timetable, or by train order, may be made, subject to prescribed signals and rules, or special instruction. Yard Switch Machine: A quick acting device electrically controlled and electrically or pneumatically operated, for positioning track switch points, and so arranged that accidental trailing of the switch points does not cause damage. Zero Sequence Component: (See Symmetrical Components).

18-101

19
ERRATA SHEET
Chapter 3, Figure 3-1: End User and Distribution Secondaries should be shown as 110 to 480V Chapter 3, Figure 3-9: Magenta line showing Neutral Current should be flat (no ripple) Chapter 4, Figure 4-24: Poor quality reproduction redraw Chapter 4, Figure 4-29: Poor quality reproduction redraw Chapter 4, Figure 4-39: Poor quality reproduction replace Chapter 4, Figure 4-75: All track resistors should be shown as variable/adjustable resistors Chapter 4, Figure 4-76: Both track resistors should be shown as variable/adjustable resistors Chapter 4, Figure 4-80: Poor quality reproduction redraw Chapter 11, Figure 11-132: Ballast Resistivity should be shown as 3 ohmkft Chapter 12, Figure 12-18: Poor quality reproduction replace Chapter 12, Figure 12-18: Exdx should be shown as E dx Chapter 12, Figure 12-38: 100-ohm m ballast should be shown as 100-ohmkft ballast 100-ohm m soil should be shown as 100-ohmm soil Chapter 12, Figure 12-39: 100-ohm m ballast should be shown as 100-ohmkft ballast 100-ohm m soil should be shown as 100-ohmm soil Chapter 12, Figure 12-40: The vertical axis may be mislabeled. It should possibly be Measured Rail Voltage (volts)

19-1

Export Control Restrictions Access to and use of EPRI Intellectual Property is granted with the specific understanding and requirement that responsibility for ensuring full compliance with all applicable U.S. and foreign export laws and regulations is being undertaken by you and your company. This includes an obligation to ensure that any individual receiving access hereunder who is not a U.S. citizen or permanent U.S. resident is permitted access under applicable U.S. and foreign export laws and regulations. In the event you are uncertain whether you or your company may lawfully obtain access to this EPRI Intellectual Property, you acknowledge that it is your obligation to consult with your companys legal counsel to determine whether this access is lawful. Although EPRI may make available on a case-by-case basis an informal assessment of the applicable U.S. export classification for specific EPRI Intellectual Property, you and your company acknowledge that this assessment is solely for informational purposes and not for reliance purposes. You and your company acknowledge that it is still the obligation of you and your company to make your own assessment of the applicable U.S. export classification and ensure compliance accordingly. You and your company understand and acknowledge your obligations to make a prompt report to EPRI and the appropriate authorities regarding any access to or use of EPRI Intellectual Property hereunder that may be in violation of applicable U.S. or foreign export laws or regulations.

The Electric Power Research Institute (EPRI) The Electric Power Research Institute (EPRI), with major locations in Palo Alto, California, and Charlotte, North Carolina, was established in 1973 as an independent, nonprofit center for public interest energy and environmental research. EPRI brings together members, participants, the Institutes scientists and engineers, and other leading experts to work collaboratively on solutions to the challenges of electric power. These solutions span nearly every area of electricity generation, delivery, and use, including health, safety, and environment. EPRIs members represent over 90% of the electricity generated in the United States. International participation represents nearly 15% of EPRIs total research, development, and demonstration program.

Together...Shaping the Future of Electricity

Program:
2006 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power Research Institute and EPRI are registered service marks of the Electric Power Research Institute, Inc.
Printed on recycled paper in the United States of America

High Power Transmission and Substation Electromagnetic Compatibility


1012652

ELECTRIC POWER RESEARCH INSTITUTE

3420 Hillview Avenue, Palo Alto, California 94304-1338 PO Box 10412, Palo Alto, California 94304-1338 USA 800.313.3774 650.855.2121 askepri@epri.com www.epri.com

S-ar putea să vă placă și